arXiv:1305.5139v5 [math.RA] 23 Feb 2015 Trust. core algebra, Azumaya , Brauer form, bilinear general C Atal,w osdrasih enmn httksit con t account into takes that refinement slight a consider we (Actually, supin r l rsm ftefloigcniin equivalent: conditions following the of some or all are assumptions ento nscin9.Let 9). section in definition nouinat-uoopimMrt qiaec;sescin2)N (A)= 2.) ously section see equivalence; involution/anti-automorphism/Morita n()ad(i.(h aewhere case (The (ii). and (i) in ncase In agba.I a hw ySlmn[9 h .]that: 3.1] Th. [19, Saltman by shown was It . -algebra 2010 Date ∗ e od n phrases. and words Key hsrsac a upre ya R rn 263 n ythe by and #226135 grant ERC an by supported was research This Let h oiainfrteqeto oe rmAuaaagba w r (we algebras Azumaya from comes question the for motivation The (A) (C) (B) -aladdress E-mail isenIsiueo ahmtc,Hbe nvriyo Jeru of University Hebrew Mathematics, of Institute Einstein i)If (ii) (i) IG HTAEMRT QIAETT THEIR TO EQUIVALENT MORITA ARE THAT RINGS oebr3 2018. 3, November : R ahmtc ujc Classification. Subject Mathematics C A R R R rtkn fadol if only and if kind first A ( etito to restriction Abstract. smtvtdb hoe fSlmnwhich for Saltman equivalent of are theorem a by motivated is uoopim (C) automorphism, oarn ihivlto,(B) involution, with a ring for a equivalent following to the of more or one ute idassumptions, mild further s tt hwta (C) that show to it use forms edmntaeta (B) that demonstrate we uhta M that such n s tt osrc xlcteape fsc algebras. such of examples explicit construct existen to the for it 2 use exponent and of algebras Azumaya test Theorem. to Saltman’s method of proof Knus-Parimala-Srinivas the on eut fsmlrflvraeas band o xml:If example: for obtained, also are flavor similar of results A ssmlcladcnetd ata losoe htoecntake can one that showed also Saltman connected, and semilocal is earn.Ti ae scnendwt h usino ne what under of question the with concerned is paper This ring. a be C/C sBae qiaett nAuaaalgebra Azumaya an to equivalent Brauer is sBae qiaett nAuaaalgebra Azumaya an to equivalent Brauer is sMrt qiaetto equivalent Morita is sMrt qiaett igwt nanti-automorphism, an with ring a involution, to with equivalent ring Morita a is to equivalent Morita is ⇒ ⊗ (B)= A f[0,w rsn eea ahnr oatc h problem the attack to machinery general a present we [10], of 0 σ ) saGli xeso ihGli group Galois with extension Galois a is : G ⇒ [email protected] n ecnie h olwn rbe:Udrwa supin a assumptions what Under problem: following the consider We sslt(..tiili h rurgopof group Brauer the in trivial (i.e. split is ( C,s n sitrse nsoig(B)= showing in interested is one so (C), R C a nivlto,te M then involution, an has ) is zmy algebras Azumaya oiaeuvlne niatmrhs,ivlto,bili involution, anti-automorphism, equivalence, Morita R σ sMrt qiaett t poiern.Teproblem The ring. opposite its to equivalent Morita is fadol ftecrsrcinagbaCor algebra corestriction the if only and if C ⇐⇒ = 6 ⇒ A C R eacmuaiern n let and ring commutative a be RY .FIRST A. URIYA A.Ormtosas ieanwperspective new a give also methods Our (A). OPPOSITES sBae qiaetto equivalent Brauer is R safil sa ale euto let ..se[1, see e.g. Albert, of result earlier an is field a is tefhsa nouin ncnrs othat, to contrast In involution. an has itself 1. B when (B) R op sMrt qiaett igwt nanti- an with ring a to equivalent Morita is Overview teopst igof ring opposite (the 13,1D0 60,16W10. 16H05, 16D90, 11E39, . ae nterecent the on Based . 1 R 2 ssmlclor semilocal is ( roughly R a nivlto,adunder and involution, an has ) ∗ R (A) : tito,smlclring. semilocal striction, ttsta l conditions all that states B B R R G iha nouinwhose involution an with sMrt equivalent Morita is A iha nouino the of involution an with sasmlclring semilocal a is Q op ial,w iea give we Finally, = R eo involutions, of ce -finite eea bilinear general ⇒ . ). C { ayDvsFellowship Davis Lady salem 0 A r(C)= or (A) 1 A ). σ , Further . ea Azumaya an be } he t htobvi- that ote ( and , σ type C/C ) then 1), 6= re erform, near cl the ecall ⇒ 0 B ( fthe of A (B). = = ) A 2 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES

Ths. 10.19 & 10.22].) Two Azumaya algebras are Brauer equivalent if and only if they are Morita equivalent as C-algebras ([3, Cor. 17.2]), so (i) can be understood as: (C)=⇒(A) for Azumaya algebras, provided the Morita equivalence is “of the first kind”. A simpler proof of Saltman’s Theorem was later found by Knus, Parimala and Srinivas ([13, §4]). In this paper, we use general bilinear forms, introduced in [10], to give partial answer to Saltman’s problem. More precisely, we show that the conditions (A) and (B) above can be phrased in terms of existence of certain bilinear forms, and use this observation to give some positive and negative results, Saltman’s Theorem in particular.

We show that (C)=⇒(B) when R is semilocal or when dimQ(R ⊗Z Q) and | ker(R → R ⊗Z Q)| are finite (i.e. when R is Q-finite). Further results of simi- lar flavor are also obtained. These include: (1) If R is semiperfect (e.g. artinian) and Morita equivalent to Rop, then S, the basic ring that is Morita equivalent to R, has an anti-automorphism. In addition, if R has an involution, then so does M2(S). (2) Suppose R is semilocal. If Mn(R) has an involution, then so does M2(R), and under mild assumptions (see Theorem 7.3 below), R itself has an in- volution.

In the special case that R is a , (2) implies that Mn(R) has an involution if and only if R has an involution, a result obtained by Albert (e.g. [1, Th. 10.12]) when [R : Cent(R)] < ∞ and by Herstein (e.g. [11, Th. 1.2.2]) in the general case. In contrast to the above, we show (B)=6 ⇒(A) even when R is a finite dimensional algebra over a field, and even when it has an anti-automorphism fixing the center pointwise. Whether (C)=⇒(B) in general is still open (and we believe this is not the case). We continue by describing the proof of Saltman’s Theorem by Knus, Parimala and Srinivas ([13, §4]) from the perspective of our methods. Namely, we recover this proof as an application of our characterization of (A) in terms of general bilinear forms. This suppresses some of the computations of [13, §4]. We also recover the sharpening of Saltman’s Theorem in case the base ring is semilocal. Finally, we apply our methods to construct non-trivial Azumaya algebras over Dedekind domains satisfying the conditions of Saltman’s Theorem, but not admitting an involution. (However, these examples are not optimal; see section 11.) The main problem of the paper was suggested to us by David Saltman himself after we gave a talk about general bilinear forms at the 10th Brauer Group confer- ence. The idea to apply the same methods for the construction of Azumaya algebras of exponent 2 not admitting an involution was communicated to us by Asher Auel, after posting a previous version of this work on the internet. We deeply thank both of them for their contribution. Sections 2 and 3 are preliminaries: Section 2 recalls the basics of Morita theory, and section 3 recalls general bilinear forms. In section 4, we give a criterion in terms of bilinear forms to when a ring is Morita equivalent to a ring with an involution (resp. anti-automorphism). This criterion is the core of this paper and it is used several times later. In sections 5–7, we show that (C)=⇒(B) under certain finiteness assumptions, as well as other results of the same flavor. Section 8 demonstrates that (B)=6 ⇒(A). The rest of the paper concerns applications to Azumaya algebras: Section 9 recalls some facts about Azumaya algebras, in section 10 we show how our methods can reproduce the proof of Saltman’s Theorem given in [13], and in section 11, we construct Azumaya algebras of exponent 2 without involution. Finally, section 12 presents some questions that we were unable to answer. RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 3

Notation and conventions: Unless specified otherwise, all rings are assumed to have a unity and ring homomorphisms are required to preserve it. Given a ring R, denote its set of invertible elements by R×, its center by Cent(R), and its Jacobson radical by Jac(R). The n×n matrices over R are denoted by Mn(R). The of right R-modules is denoted by Mod-R and the category of f.g. projective right R-modules is denoted by proj-R. For a subset X ⊆ R, we let CentR(X) denote the centralizer of X in R. If a M can be viewed as a module over several rings, we use MR (resp. RM) to denote “M, considered as a right (resp. left) R-module”. Endomorphisms of left (right) modules are applied on the right (left). Throughout, a semisimple ring means a semisimple artinian ring.

2. Morita Theory In this section, we recall some facts from Morita Theory, and also give a refine- ment of the main problem of the paper. See [15, §18] or [18, §4.1] for proofs and further details.

2.1. Morita Theory. Let R be a ring. A right R-module M is called a generator if every right R-module is an epimorphic image of i∈I M for I sufficiently large, n or equivalently, if RR is a summand of M for some n ∈ N. The module M is called a progenerator if M is a generator, finitely generatedL and projective. In this case, we also call M a (right) R-progenerator. Let S be another ring. An (S, R)-progenerator is an (S, R)- P such that PR is a progenerator and S = End(PR) (i.e. every endomorphism of PR is of the form p 7→ sp for unique s ∈ S). In this case, SP is also a progenerator and R = End(SP ).

The rings R, S are said to be Morita equivalent, denoted R ∼Mor S, if the categories Mod-R and Mod-S are equivalent. Morita’s Theorems assert that: (1) For all equivalences1 F : Mod-S → Mod-R there exists an (S, R)-progenera- tor P such that FM is naturally isomorphic to M ⊗S P for all M ∈ Mod-S. (2) Conversely, for any (S, R)-progenerator P the functor (−)⊗S P : Mod-S → Mod-R is an equivalence of categories. (3) There is a one-to-one correspondence between equivalences of categories F : Mod-S → Mod-R (considered up to natural ) and iso- morphism classes of (S, R)-progenerators. The correspondence maps the composition of two equivalences to the tensor product of the corresponding progenerators. Note that (3) allows us to “divide” progenerators. For example, if P is an (S, R)- progenerator and Q is an (S′, R)-progenerator, then there exists an (S′,S)-pro- generator V , unique up to isomorphism, such that Q =∼ V ⊗S P . Indeed, if F : Mod-S → Mod-R and G : Mod-S′ → Mod-R correspond to P and Q, respectively, then V is the progenerator corresponding to F −1G : Mod-S′ → Mod-S. (In fact, one can take V = HomR(P,Q).)

Every (S, R)-progenerator P induces an isomorphism σP : Cent(R) → Cent(S) given by σP (r) = s where s is the unique element of Cent(S) satisfying sp = pr for all p ∈ P . As σP depends only on the isomorphism class of P , it follows that any equivalence of categories F : Mod-S → Mod-R induces an isomorphism σF : Cent(R) → Cent(S).

1 According to textbooks, an equivalence between two categories A and B consists of a quartet (F,G,δ,ε) such that F : A → B and G : B → A are functors and δ : idA → GF and ε : idB → FG are natural . We do not need this detailed description here and hence we only specify F . In this case, the implicit functor G is determined up to natural isomorphism. 4 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES

Let C be a and assume R and S are C-algebras. We say that R and S are Morita equivalent as C-algebras or over C, denoted R ∼Mor/C S, if there exists an equivalence F : Mod-S → Mod-R such that σF (c · 1R) = c · 1S for all c ∈ C. Equivalently, this means that there exists an (S, R)-progenerator P such that cp = pc for all p ∈ P and c ∈ C. If S is an arbitrary ring that is Morita equivalent to R and F : Mod-S → Mod-R is any equivalence, then we can make S into a C-algebra by letting C act on S via σF . In this setting, we have S ∼Mor/C R. 2.2. Types. To make the phrasing of some results in the next sections easier, we now introduce types: Let C be a commutative ring and let R and S be central C-algebras. (The algebra R is central if the map c 7→ c · 1R : C → Cent(R) is an isomorphism.) Every equivalence of categories F : Mod-S → Mod-R gives rise to an isomorphism σF : Cent(R) → Cent(S). As both Cent(R) and Cent(S) are isomorphic to C, we can realize σF as an automorphism of C, which we call the type of F . (For example, when F is of type idC , R is Morita equivalent to S as C-algebras.) Likewise, the type of an (S, R)-progenerator P is the type of the equivalence induced by P . Namely, it is the unique automorphism σ of C satisfying σ(c)p = pc for all p ∈ P , c ∈ C. Let α be an automorphism or an anti-automorphism of R. The type of α is defined to be its restriction to C = Cent(R). For example, an involution of R is of the first kind (i.e. it fixes Cent(R) pointwise) if and only if its type is idC . We now make an essential sharpening of the problem presented in section 1. Let R be a ring, let C = Cent(R) and let σ ∈ Aut(C). We look for sufficient conditions ensuring that some or all of the following are equivalent: (A) R is Morita equivalent over C to a (necessarily central) C-algebra with involution of type σ. (B) R is Morita equivalent over C to a (necessarily central) C-algebra with an anti-automorphism of type σ. (C) R is Morita equivalent to Rop via equivalence of type σ (Rop is considered as a C-algebra in the obvious way). Again, (A)=⇒(B)=⇒(C),2 so we want to show that (B)=⇒(A) or (C)=⇒(B). Satl- man’s Theorem ([19, Th. 3.1]) for involutions of the first kind can now be phrased as (C)=⇒(A) when R/C is Azumaya and σ = idC . 2.3. Progenerators and Scalar Extension. We proceed by recalling several facts about the behavior of progenerators with respect to scalar extension. Throughout, C is a commutative ring and R is a C-algebra. All tensor products are taken over C. If σ is an automorphism of C, then Rσ denotes the C-algebra obtained from R by letting C act via σ. Observe that for all M,N ∈ Mod-R, HomR(M,N) admits a (right) C-module structure given by (fc)m = (fm)c (f ∈ ′ ′ HomR(M,N), c ∈ C, m ∈ M). In addition, if R is another C-algebra and M ∈ Mod-R, then M ⊗M ′ has a right R⊗R′-module structure given by (m⊗m′)(r⊗r′)= mr ⊗ m′r′. We start with recalling a well-known fact: Proposition 2.1. Let R′ be a C-algebra and let X, Y ∈ Mod-R, X′, Y ′ ∈ Mod-R′. Then there is a natural homomorphism of C-modules ′ ′ ′ ′ HomR(X, Y ) ⊗ HomR′ (X , Y ) → HomR⊗R′ (X ⊗ X , Y ⊗ Y )

2 To see that (B)=⇒(C), let S be a central C-algebra admitting an anti-automorphism α of ′ type σ, and assume that there is an (S, R)-progenerator P of type idC (i.e. S ∼Mor/C R). Let P be the (Rop,S)-bimodule obtained from P by setting rop · p · s = sαpr. Then P ′ is an (Rop,S)- ′ op progenerator, hence P ⊗S P is an (R , R)-progenerator, and the latter is easily seen to have type σ. RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 5 given by setting (f ⊗f ′)(x⊗x′)= fx⊗f ′x′. This homomorphism is an isomorphism ′ whenever XR, XR′ are finitely generated projective. Proof. See [8, pp. 14], for instance. Alternatively, one can easily check that the map ′ ′ in the proposition is an isomorphism in case X = RR and X = RR′ . Since the map is natural (in the functorial sense), it is additive in X and X′, hence it is an isomorphism whenever X and X′ are summands of Rn and R′m, respectively.  As a special case, we get:

Proposition 2.2. Let S be a C-algebra, let RS := R⊗S, and set XS = X⊗S for all ′ ′ X ∈ Mod-R. (Observe that XS is a right RS -module via (x ⊗ s)(r ⊗ s )= xr ⊗ ss .) Then for all X, Y ∈ Mod-R, there is a natural homomorphism

HomR(X, Y ) ⊗ S → HomRS (XS , YS) given by setting (f ⊗ s)(x ⊗ s′) = fx ⊗ ss′ for s,s′ ∈ S, x ∈ X, r ∈ R. This homomorphism is an isomorphism when XR is finitely generated projective. Proof. Take R′ = X′ = Y ′ = S in Proposition 2.1.  Proposition 2.3. Let S and D be C-algebras and let P be an (S, R)-progenerator σ σ of type σ ∈ Aut(C). Put RD = R ⊗ D, SD = (S ) ⊗ D and PD = P ⊗ D, and σ ′ ′′ endow PD with an (SD, RD)-bimodule structure by setting (s ⊗ d)(p ⊗ d )(r ⊗ d )= ′ ′′ σ (spr) ⊗ (dd d ). Then PD is an (SD, RD)-progenerator of type idD. σ Proof. Since P has type σ, End(PR) =∼ S as C-algebras. By Proposition 2.2, ∼ σ σ EndRD (PD) = HomRD (PD, PD) = HomR(P, P ) ⊗ D = S ⊗ D = SD. It is routine σ to verify that the action of SD on PD via endomorphisms is the action specified in the proposition. Finally, that P is a right R-progenerator implies that PD is a right σ RD-progenerator, hence PD is an (SD, RD)-progenerator. The type is verified by straightforward computation.  Proposition 2.4. Let P be an (S, R)-progenerator and let N and M be the prime radicals (resp. Jacobson radicals) of R and S, respectively. View P := P/PN as a right R := R/N-module. Then PN = MP , hence P = P/MP admits a left S-module structure. Furthermore, P is an (S, R)-progenerator. Proof. By [15, Pr. 18.44], there is an isomorphism between the lattice of R-ideals and the lattice of (S, R)-submodules of P given by I 7→ P I. Similarly, the ideals of S correspond to (S, R)-submodules of P via J 7→ JP , hence every IER admits a unique ideal J E S such that JP = P I. The ideal J can also be described as HomR(P,PI). This description implies that S/J = HomR(P, P )/ Hom(P,PI) =∼ ∼ HomR(P,P/PI) = EndR/I (P/PI). Thus, P/PI is an (S/J, R/I)-progenerator. Choose I = N. Then by [15, Cor. 18.45] (resp. [15, Cor. 18.50]) J = M, so we are done. 

3. General Bilinear Forms General bilinear forms were introduced in [10]. In this section, we recall their basics and record several facts to be needed later. When not specified, proofs can be found in [10, §2]. Throughout, R is a (possibly non-commutative) ring. Definition 3.1. A (right) double R-module is an additive group K together with two operations ⊙0, ⊙1 : K × R → K such that K is a right R-module with respect to each of ⊙0, ⊙1 and

(k ⊙0 a) ⊙1 b = (k ⊙1 b) ⊙0 a ∀ k ∈ K, a,b ∈ R.

We let Ki denote the R-module obtained by letting R act on K via ⊙i. 6 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES

′ ′ ′ For two double R-modules K,K , we define Hom(K,K ) = HomR(K0,K0) ∩ ′ HomR(K1,K1). This makes the class of double R-modules into an abelian cate- op gory (which is isomorphic to Mod-(R ⊗Z R) and also to the category of (R , R)- ).3 An involution of a double R-module K is an additive bijective map θ : K → K satisfying θ ◦ θ = idK and θ θ (k ⊙i a) = k ⊙1−i a ∀a ∈ R, k ∈ K, i ∈{0, 1} . Definition 3.2. A (general) bilinear space over R is a triplet (M,b,K) such that M ∈ Mod-R, K is a double R-module and b : M × M → K is a biadditive map satisfying

b(xr, y)= b(x, y) ⊙0 r and b(x, yr)= b(x, y) ⊙1 r for all x, y ∈ M and r ∈ R. In this case, b is called a (general) bilinear form (over R). Let θ be an involution of K. The form b is called θ-symmetric if b(x, y)= b(y, x)θ ∀x, y ∈ M. See [10, §2] for various examples of general bilinear forms. Fix a double R-module K and let i ∈{0, 1}. The i-K-dual (or just i-dual) of an R-module M is defined by [i] M := HomR(M,K1−i) . [i] Note that M admits a right R-module structure given by (fr)(m) = (fm) ⊙i r (where f ∈ M [i], r ∈ R and m ∈ M). In fact, M 7→ M [i] is a left-exact contravariant [i] functor from Mod-R to itself, which we denote by [i]. Also observe that R =∼ Ki via f ↔ f(1). Let b : M × M → K be a (general) bilinear form. The left adjoint map and right adjoint map of b are defined as follows: ℓ [0] ℓ Adb : M → M , (Adbx)(y)= b(x, y) ,

r [1] r Adb : M → M , (Adb x)(y)= b(y, x) , ℓ r where x, y ∈ M. It is straightforward to check that Adb and Adb are right R-linear. r ℓ We say that b is right (resp. left) regular if Adb (resp. Adb) is bijective. If b is both right and left regular, we say that b is regular. Left and right regularity are not equivalent properties; see [10, Ex. 2.6]. Assume b is regular. Then for every w ∈ EndR(M) there exists a unique element α w ∈ EndR(M) such that b(wx,y)= b(x, wαy) ∀x, y ∈ M.

α The map w 7→ w , denoted α, turns out to be an anti-automorphism of EndR(M) which is called the (right) corresponding anti-automorphism of b. (The left corre- sponding anti-automorphism of b is the inverse of α.) If b is θ-symmetric for some involution θ : K → K, then α is easily seen to be an involution. We say that two bilinear spaces (M,b,K), (M,b′,K′) are similar if there is an isomorphism f : K → K′ such that b′ = f ◦ b. It is easy to see that in this case, b and b′ have the same corresponding anti-automorphism, provided they are regular.

3 op The usage of double R-modules, rather than (R , R)-modules or R ⊗Z R-modules, was more convenient in [10], so we follow the notation of that paper. In addition, the notion of double module is more natural when considering bilinear forms as a special case of multilinear forms, where the form takes values in a (right) multi-R-module. RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 7

Theorem 3.3 ([10, Th. 5.7]). Let M be a right R-generator. Then the map sending a regular bilinear form on M to its corresponding anti-automorphism gives rise to a one-to-one correspondence between the class of regular bilinear forms on M, con- sidered up to similarity, and the anti-automorphisms of EndR(M). Furthermore, the correspondence maps the equivalence classes of forms which are θ-symmetric with respect to some θ to the involutions of EndR(M).

The theorem implies that every anti-automorphism α of EndR(M), with M an R-generator, is induced by some regular bilinear form on M, which is unique up to similarity. We will denote this form by bα and the double R-module in which it takes values by Kα. In case α is an involution, then bα is symmetric with respect to some involution of Kα, which we denote by θα. The objects Kα, bα and θα can be explicitly constructed as follows: Let W = EndR(M). Then M is a left W -module. Using α, we may view M as a right W - module by defining m · w = wαm. Denote by M α the right W -module obtained. Then Kα = Mα ⊗W M. We make Kα into a double R-module by defining (x ⊗ y)⊙0 r = x⊗yr and (x⊗y)⊙1 r = xr ⊗y. The form bα is given by bα(x, y)= y ⊗x, θα and when α is an involution, θα : Kα → Kα is defined by (x ⊗ y) = y ⊗ x. We will also need the following technical proposition. Proposition 3.4 ([10, Lm. 7.7]). Fix a double R-module K. For M ∈ Mod-R, [1][0] [1] define ΦM : M → M by (ΦM x)f = f(x) for all x ∈ M and f ∈ M . Then:

(i) {ΦM }M∈Mod-R is a natural transformation from idMod-R to [0][1] (i.e. for ′ ′ [1][0] all N,N ∈ Mod-R and f ∈ HomR(N,N ), one has f ◦ ΦN =ΦN ′ ◦ f). ′ (ii) Φ is additive (i.e. ΦN⊕N ′ =ΦN ⊕ ΦN ′ for all N,N ∈ Mod-R). r [0] ℓ (iii) (Adb ) ◦ ΦM = Adb for every general bilinear form b : M × M → K. [1][0] (iv) R can be identified with EndR(K1). Under that identification, (ΦRr)k = k ⊙0 r for all r ∈ R and k ∈ K. To finish, we recall that the orthogonal sum of two bilinear spaces (M,b,K) and (M ′,b′,K) is defined to be (M ⊕ M ′,b ⊥ b′,K) where (b ⊥ b′)(x ⊕ x′,y ⊕ y′) = b(x, y)+ b′(x′,y′). The form b ⊥ b′ is right regular if and only if b and b′ are right regular.

4. Double Progenerators The observation which forms the basis of all results of this paper is the fact that whether a ring R is equivalent to a ring with an anti-automorphism (resp. involu- tion) can be phrased in terms of existence of certain bilinear forms (resp. double R-modules). In this section, we state and prove this criterion (Proposition 4.3 and Theorem 4.5). Throughout, R is a ring and C = Cent(R). Recall from section 2 that for all M,N ∈ Mod-R, HomR(M,N) admits a (right) C-module structure. In particular, EndR(M) is a C-algebra. Let K be a double R-module. Then K can be viewed as an (Rop, R)-bimodule op op by setting a · k · b = k ⊙0 b ⊙1 a. If this bimodule is an (R , R)-progenerator, we say that K is a double R-progenerator. The type of a double R-progenerator is the type of its corresponding (Rop, R)-module. Namely, it is the automorphism σ ∈ Aut(C) satisfying k ⊙0 c = k ⊙1 σ(c) for all c ∈ Cent(R), k ∈ K. It turns out that Theorem 3.3 is useful for producing double progenerators. Lemma 4.1. Let M be an R-progenerator and let α be an anti-automorphism of EndR(M). Then Kα is a double R-progenerator. Viewing EndR(M) as a C-algebra, the type of Kα is the type of α. 8 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES

α Proof. Let W = EndR(M). Recall from section 3 that Kα = M ⊗W M. Since α op MR is an R-progenerator, M is a (W, R)-progenerator. Endow M with a left R - module structure by setting ropm = mr. Then M α is an (Rop, W )-bimodule and, moreover, it is easily seen to be an (Rop, W )-progenerator (because M is a (W, R)- progenerator). By Morita theory (see section 2), this means that Kα = Mα ⊗W M op is an (R , R)-progenerator, as required. That the type of Kα is the type of α follows by straightforward computation.  Lemma 4.2. Fix a double R-progenerator K. Then:

(i) ΦM is an isomorphism for all M ∈ proj-R (see Proposition 3.4 for the definition of Φ). (ii) The (contravariant) functors [0], [1] : proj-R → proj-R are mutual inverses. (iii) Let (M,b,K) be a bilinear space with M ∈ proj-R. Then b is right regular if and only if b is left regular.

Proof. (i) Since K is a double R-progenerator, every automorphism of K1 is of the form k 7→ k ⊙0 r for a unique r ∈ R. Hence, ΦR is an isomorphism by Proposition 3.4(iv). By the additivity of Φ (Proposition 3.4(ii)), ΦRn is also an isomorphism for all n. As every M ∈ proj-R admits an M ′ ∈ proj-R with M ⊕M ′ =∼ n R , it follows that ΦM is an isomorphism (since ΦM ⊕ ΦM ′ =ΦRn ). [i] (ii) Recall that R =∼ Ki ∈ proj-R. Since the functor [i] is additive, it follows that [i] takes proj-R into itself. By (i), Φ : idproj-R → [1][0] is a natural isomorphism. [0][1] [0] For all M ∈ Mod-R, define ΨM : M → M by (ΨM x)f = f(x), where f ∈ M and x ∈ M. Then a similar argument shows that Ψ : idproj-R → [0][1] is a natural isomorphism. It follows that [0] and [1] are mutual inverses. r [0] ℓ r (iii) By Proposition 3.4(iii), (Adb) ◦ ΦM = Adb, so by (i), that Adb is bijective ℓ implies Adb is bijective. The other direction follows by symmetry. (Use the identity ℓ [1] r (Adb) ◦ ΨM = Adb .)  Proposition 4.3. The ring R is Morita equivalent over C to a (necessarily central) C-algebra with an anti-automorphism of type σ ∈ Aut(C) if and only if there exists a regular bilinear space (M,b,K) such that M is an R-progenerator and K is a double R-progenerator of type σ. Proof. If (M,b,K) is a regular bilinear space, we may associated to b its correspond- ing anti-automorphism α : EndR(M) → EndR(M). Since M is an R-progenerator, EndR(M) ∼Mor/C R (recall that we view EndR(M) as a C-algebra). In addition, for all x, y ∈ M and c ∈ C, we have b(x, y) ⊙1 σ(c) = b(x, y) ⊙0 c = b(cx,y) = α α b(x, c y)= b(x, y) ⊙1 c , so since K1 is faithful (because it is a progenerator), the type of α is σ. Conversely, let M be an R-progenerator and assume EndR(M) has an anti- automorphism of type σ. Then by Theorem 3.3, bα : M × M → Kα is a regular bilinear form and Kα is a double R-progenerator of type σ by Lemma 4.1.  Lemma 4.4. Let K be a double R-module with involution θ. For every M ∈ [0] [1] Mod-R, define uθ,M : M → M by uθ,M (f) = θ ◦ f. Then uθ,M is a natural isomorphism of right R-modules. Proof. By computation.  The proof of following theorem demonstrates an idea that we wish to stress: One can construct involutions on rings that are Morita equivalent to R by constructing symmetric general bilinear forms. Note that all possible involutions are obtained in this manner by Theorem 3.3. (However, this fails if we limit ourselves to standard sesquilinear forms. Indeed, the base ring may not even have an anti-automorphism; see [10, Exs. 2.4 & 2.7], or section 11.) RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 9

Theorem 4.5. The ring R is Morita equivalent over C to a C-algebra with an involution of type σ ∈ Aut(C) if and only if there exists a double R-progenerator of type σ admitting an involution.

Proof. Assume MR is an R-progenerator such that EndR(M) has an involution of type σ. Then, as in the proof of Proposition 4.3, Kα is a double R-progenerator of type σ. Since α is an involution, Kα has an involution, namely θα. Conversely, assume K is a double R-progenerator of type σ with an involution θ. Let P be any right R-progenerator and let M = P ⊕ P [1]. Define b : M × M → K by b(x ⊕ f,y ⊕ g)= gx + (fy)θ. The form b is clearly θ-symmetric. We claim that b is regular. By Lemma 4.2(iii) (or since b is θ-symmetric), it is enough to show only right regularity. Indeed, a straightforward computation shows that

r 0 idP [1] [1] [1] [1][1] Adb = ∈ HomR(P ⊕ P , P ⊕ P ) . u [1] ◦ Φ 0  θ,P P  As ΦP and uθ,P [1] are both bijective by Lemmas 4.2(i) and 4.4, respectively, so r is Adb . Let α : EndR(M) → EndR(M) be the anti-endomorphism associated with b. Then α is an involution (since b is θ-symmetric), and, as in the proof of Proposition 4.3, α is of type σ.  Using very similar ideas, one can show the following variation, which states when a particular ring S that is Morita equivalent to R over C has an anti-automorphism or an involution of a given type. Compare with [19, Th. 4.2].

Proposition 4.6. Let M be a right R-progenerator. Then EndR(M) has an anti- automorphism (involution) if and only if there exists a double R-progenerator K of type σ (admitting an involution θ) and a right regular (θ-symmetric) bilinear form b : M × M → K.

5. Rings That Are Morita Equivalent to Their Opposites Let R be a ring. In this section, we use Proposition 4.3 to show that under op certain finiteness assumptions, R ∼Mor R implies that R is Morita equivalent to a ring with an anti-automorphism (i.e. (C)=⇒(B)). Whether this holds in general is still open. Henceforth, we freely consider (Rop, R)-progenerators as double R- progenerators and vice versa (see section 4). We begin with two lemmas whose purpose is to show that the functors [0] and [1] commute with certain scalar extensions. Lemma 5.1. Let K be an (Rop, R)-progenerator of type σ ∈ Aut(Cent(R)), let C ⊆ Cent(R) be a subring fixed pointwise by σ, and let D/C be a commutative ring extension. For every P ∈ proj-R, let PD := P ⊗C D ∈ proj-RD. Make KD into an op op (RD , RD)-module as in Proposition 2.3. Then KD is an (RD , RD)-progenerator [1] [1] and (P )D =∼ (PD) for all P ∈ proj-R (the functor [1] is computed with respect to K in the left hand side and with respect to KD in the right hand side). op Proof. That KD is an (RD , RD)-progenerator follows from Proposition 2.3 (observe op σ op op [1] ∼ [1] that (R )D = (R )D = (RD) since σ fixes C). That (P )D = (PD) follows from Proposition 2.2.  Lemma 5.2. Let K be an (Rop, R)-progenerator and let N be the prime (resp. Jacobson) radical of R. For all P ∈ proj-R, define P := P/PN =∼ P ⊗R (R/N) ∈ ∼ op op [1] ∼ [1] proj-R/N. Then EndR(K) = R , K is an (R , R)-progenerator and P = P (the functor [1] is computed with respect to K in the left hand side and with respect to K in the right hand side). 10 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES

op Proof. By Proposition 2.4, KN = N opK and K is an (R , R)-progenerator. Let [1] P ∈ proj-R. We claim that P N = HomR(P, (KN)0) (we consider KN as a double [1] R-module here). Indeed, by definition, P N = HomR(P,K0)N ⊆ HomR(P, (K ⊙1 op N)0) = HomR(P, (N K)0) = HomR(P, (KN)0), so there is a natural inclusion [1] P N ⊆ HomR(P, (KN)0). As this inclusion is additive in P , it is enough to verify the equality for P = RR, which is routine. Now, we get Hom (P,K ) Hom (P,K ) [1] [1] R 0 R 0 ∼ ∼ P = [1] = = HomR(P, K0) = HomR(P, K0)= P , P N HomR(P, (KN)0) as required. 

Recall that a ring R is called Q-finite if both dimQ(R ⊗Z Q) and the cardinality of ker(R → R ⊗Z Q) are finite. The following lemma is inspired by [3, Pr. 18.2]. Lemma 5.3. Assume R is Q-finite and let N be the prime radical of R. Then N is nilpotent and R/N =∼ T × Λ where T is a semisimple finite ring and Λ is a subring of a semisimple Q-algebra E such that ΛQ = E.

Proof. Let T = ker(R → R ⊗Z Q). Then T is an ideal of R. Consider T as a non- unital ring and let J = Jac(T ) (i.e. J is the intersection of the kernels of all right T -module morphisms T → M where M is a simple right T -module with MT = M). Arguing as in [3, Pr. 18.2], we see that J is also an R-ideal. Since J and all its submodules are finite (as sets), J n = 0 for some n (because MJ ⊆ N for any right T -module M and any maximal submodule N ≤ M). In particular, J ⊆ N. Replacing R with R/J, we may assume J = 0. Now, T is semisimple and of finite length, hence it has a unit e. As er, re ∈ T for all r ∈ R, we see that er = ere = re. Thus, e ∈ Cent(R) and R =∼ T × (1 − e)R. As Λ := (1 − e)R is torsion-free, it is a subring of E := Λ ⊗Z Q, which is a f.d. Q-algebra by assumption. Let I = Jac(E) ∩ Λ. Then I is nilpotent, hence 0 × I ⊆ N. Replacing R by R/(0 × I), we may assume E semisimple. We are thus finished if we show that the prime radical of Λ, denoted N ′, is 0 (because then N = 0 × N ′ = 0). Indeed, by [5, Th. 2.5], Λ is noetherian (here we need E to be semisimple). Thus, N ′ is nil, hence so is N ′Q E E. But E is semisimple, so we must have N ′ ⊆ N ′Q = 0.  Theorem 5.4. Assume R is semilocal or Q-finite and let K be an (Rop, R)- n progenerator. Then for every P ∈ proj-R there is n ∈ N such that P =∼ P [1] := P [1][1]...[1] (n times). When R is semilocal, n is independent of P . Proof. Let N denote the Jacobson radical of R in case R is semilocal and the prime radical of R otherwise. In the latter case, N is nilpotent by Lemma 5.3, so N ⊆ Jac(R) in both cases. Using the notation of Lemma 5.2, observe that every P ∈ proj-R is the projective cover of P = P/PN. As projective covers are unique up to isomorphism, we have P =∼ Q ⇐⇒ P =∼ Q for all P,Q ∈ proj-R. Therefore, using Lemma 5.2, we may assume N = 0. Thus, R is semisimple or R = T × Λ with Λ ⊗Z Q semisimple and T finite and semisimple. Assume R is semisimple and let V1,...,Vt be a complete set of simple R-modules up to isomorphism. By Lemma 4.2(ii), [1] permutes isomorphism classes of f.g. projective R-modules, and since [1] is additive, it permutes the indecomposable projective modules, namely, V1,...,Vt. Therefore, there is n ∈ N (say, n = t!) such n [1] that V =∼ Vi for all i. As any P ∈ proj-R is a direct sum of simple modules, we i n get P [1] =∼ P . We also record that length(P [1]) = length(P ). Now assume R = T × Λ as above, and set E := Λ ⊗Z Q and S := T × E. Let σ ∈ Aut(Cent(R)) be the type of K and let e = (1T , 0) ∈ R. Then e is the maximal torsion idempotent in Cent(R), hence σ(e)= e. Define C := Ze+Z(1−e) RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 11 and D := Ze + Q(1 − e)+ ⊆ T × E = S. Then σ|C = idC and S =∼ R ⊗C D. ′ ′ op Let K = K ⊗C D. Then by Lemma 5.1, K is an (S ,S)-progenerator and [1] [1] (PS ) =∼ (P )S for all P ∈ proj-R (where PS := P ⊗R S). For every P ∈ proj-R, define j(P ) = length(PS). Since S is semisimple, the previous paragraphs imply that j(P [1])= j(P ). Therefore, we are done if we show that for all m ∈ N there are finitely many isomorphism classes of modules P with j(P )= m (for this implies that action of [1] have finite orbits on proj-R/=∼, which yields the theorem). Indeed, thanks to [5, Th. 2.8], up to isomorphism, there are finitely many Λ-modules of any given Z-rank, and this is easily seen to imply that there are finitely many isomorphism classes of modules P ∈ proj-R with PS of a given length, as required. 

Remark 5.5. It is also possible to prove the semilocal case of Theorem 5.4 by using a result of Facchini and Herbera [9, Cor. 2.13] stating that every semilocal ring R has only finitely many indecomposable projective modules up to isomorphism, and every f.g. projective module is a direct sum of these modules.

Corollary 5.6. Let C = Cent(R), σ ∈ Aut(C) and assume R is semilocal or Q-finite. Then R is Morita equivalent over C to a (central) C-algebra with an anti- automorphism of type σ if and only if R is Morita equivalent to Rop (as rings) via Morita equivalence of type σ.

Proof. We only check the nontrivial direction. Assume R is Morita equivalent to Rop via Morita equivalence of type σ. Then there exists a double R-progenerator n [1] K of type σ. By Theorem 5.4, there exists n ∈ N such that R =∼ RR. Let m n−1 [1] [1] M = m=0 R . Then there is an isomorphism f : M → M . This isomorphism r gives rise to a right regular bilinear space (M,b,K) with f = Adb , namely, b : M × ML → K is given by b(x, y) = (fy)x. By Lemma 4.2(iii), b is regular, so we are done by Proposition 4.3. 

The proof of Corollary 5.6 cannot be applied to arbitrary rings since there are double R-modules K for which M [1] ≇ M for all 0 6= M ∈ proj-R: Example 5.7. ⊗n Let F be a field and let R = lim−→ {M2(F ) }n∈N. Then any f.g. projective right R-module is obtained by scalar extension from a f.g. projective ⊗n module over M2(F ) ֒→ R. Using this, it not hard (but tedious) to show that ∼ 1 the (proj-R/ =, ⊕) is isomorphic to (Z[ 2 ] ∩ [0, ∞), +). If Vn is the unique ⊗n −n simple projective right module over M2(F ) , then Vn ⊗ R is mapped to 2 . Let T denote the transpose involution on M (F ). Then T = lim {T ⊗n} is an 2 −→ n∈N 2 involution of R. Let K = R ∈ proj-R. Then EndR(K) =∼ M2(R) =∼ R and using T , op b op we can identify EndR(K) with R , thus making P into an (R , R)-progenerator. We claim that for M ∈ proj-R and n ∈ N, M [1] =∼ M implies M = 0. To see this,b ∼ 1 let ϕ1 be the map obtained from [1] by identifying proj-R/ = with Z[ 2 ] ∩ [0, ∞). 2 [1] ∼ 2 ∼ 1 Then ϕ1(2) = 1 because (RR) = K1 = RR. Therefore, ϕ1(x) = 2 x for all 1 1 x ∈ Z[ 2 ] ∩ [0, ∞), which means that ϕ1(x) 6= x for all 0 6= x ∈ Z[ 2 ] ∩ [0, ∞).

6. Semiperfect Rings Let R be a semilocal ring that is Morita equivalent to its opposite ring. While Corollary 5.6 implies that R is Morita equivalent to a ring with an anti-automorphism, it does not provide any information about this ring. However, when R is semiper- fect, we can actually point out a specific ring which is Morita equivalent to R and has an anti-automorphism. 12 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES

Recall that a ring R is semiperfect if R is semilocal and idempotents lift modulo Jac(R) (e.g. if Jac(R) is nil). In this case, the map P 7→ P/P Jac(R) : proj-R/=∼ → proj-(R/ Jac(R))/=∼ is bijective (e.g. see [18, §2.9] or [4, Th. 2.1]). Thus, up to isomorphism, R admits finitely many indecomposable projective right R-modules P1,...,Pt and any P ∈ ∼ t ni proj-R can be written as P = i=1 Pi with n1,...,nt uniquely determined. In particular, R =∼ t P mi for some (necessarily positive) m ,...,m . The ring R i=1 i L 1 t R is called basic if m1 = ··· = mt = 1, namely, if RR is a sum of non-isomorphic indecomposable projectiveL modules. It is well-known that every semiperfect ring R admits a basic ring S that is Morita equivalent to it over Cent(R), and S is well-determined up to isomorphism as a Cent(R)-algebra. (Indeed, S and the map Cent(R) → S can be determined from Mod-R; see [15, Prp. 18.37], the preceding discussion, and [15, Rm. 18.43]. Once S is fixed, the Morita equivalence Mod-S → Mod-R is not unique in general.) Explicitly, we can take S = EndR(M), where M = P1 ⊕···⊕ Pt. (For example, if R = Mn(L) with L a local ring, then S = L.) Assume now that there is an (Rop, R)-progenerator K of type σ ∈ Aut(Cent(R)). Then the functor [1] must permute the isomorphism classes of P1,...,Pt (because they are the only indecomposable modules in proj-R) and hence stabilize M = P1 ⊕···⊕ Pt. Therefore, as in the proof of Corollary 5.6, EndR(M), the basic ring which is Morita equivalent to R, has an anti-automorphism of type σ. We have thus obtained the following proposition. Proposition 6.1. Let R be a semiperfect ring and let S be a basic ring that is Morita equivalent to R over Cent(R). Then R is Morita equivalent to Rop via an equivalence of type σ ∈ Aut(Cent(R)) if and only if S has an anti-automorphism of type σ. Proposition 6.1 has a version for involutions in which the ring S is replaced with M2(S). Proposition 6.2. Let R be a semiperfect ring and let S be a basic ring that is Morita equivalent to R over Cent(R). If R has an involution of type σ, then so does M2(S). Proof. Let α be an involution of R of type σ, and let K be the double R-module α obtained from R by setting k ⊙0 r = r k and k ⊙1 r = kr (k, r ∈ R). Then K is double R-progenerator of type σ admitting an involution θ := α. For any P ∈ proj-R, let bP denote the bilinear form b constructed in the proof of Theorem 4.5. [1] Then (bP , P ⊕ P ,K) is an α-symmetric regular bilinear space. Let P1,...,Pt be a complete list of indecomposable projective right R-modules up to isomorphism.

Then b := bP1 ⊥ ··· ⊥ bPt is a regular α-symmetric bilinear form defined over [1] [1] M ⊕ M , where M = P1 ⊕···⊕ Pt. Therefore, EndR(M ⊕ M ) has an involution of type σ, namely, the corresponding anti-automorphism of b. However, we have [1] [1] seen above that M =∼ M , so EndR(M ⊕ M ) =∼ M2(EndR(M)) =∼ M2(S) as Cent(R)-algebras. 

In fact, in many cases, the assumptions of Proposition 6.2 imply that S itself has an involution. The general statement and its proof will be given in the next section. However, in case S is a division ring or local with with 2 ∈ S×, there is a significantly simpler proof, with which we finish this section.

× Proposition 6.3. Let L be a local ring and let R = Mn(L). Assume 2 ∈ L or L is a division ring. Then R has an involution of type σ if and only if L has an involution of type σ. RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 13

Proof. That R has an involution when L has an involution is obvious. Conversely, let α be an involution of R and let K, θ be as in the proof of Proposition 6.2. Let P be the unique indecomposable projective right R-module. Then necessarily P =∼ P [1]. Fix an isomorphism f : P → P [1] and observe that the bilinear form b(x, y) := (fy)x+((fx)y)θ (resp. b′(x, y) := (fy)x−((fx)y)θ) is θ-symmetric (resp. r r (−θ)-symmetric). In addition, Adb + Adb′ =2f. × −1 r −1 r Assume 2 ∈ L . Then f ◦Adb +f ◦Adb′ = 2 idP . Since the r.h.s. is invertible ∼ −1 r −1 r and lies in the local ring EndR(P ) = L, one of f ◦ Adb , f ◦ Adb′ must be ′ invertible, hence one of b, b is right regular. In any case, we get that L =∼ EndR(P ) has an involution of type σ, namely, the corresponding anti-automorphism of b or b′. ′ r r When L is a division ring, P is simple, so b is regular if Adb′ 6=0. If Adb′ = 0, then b′ = 0, hence (fy)x = ((fx)y)θ for all x, y ∈ P . This means that the bilinear ′′ r ′′ form b (x, y) := (fy)x is θ-symmetric. As Adb′′ = f, b is regular.  Remark 6.4. In case L is a division ring, Proposition 6.3 follows from [11, Th. 1.2.2]. The case where L is also finite dimensional over its center was noted earlier by Al- bert (e.g. see [1, Th. 10.12]).

7. Transferring Involutions Motivated by Proposition 6.2, this section concerns the question of whether the fact that the ring Mn(R) has an involution implies that R has an involution. We shall provide a positive answer for “most” semilocal rings, but the question is still open for general semilocal rings, and even for finite dimensional algebras over fields. For non-semilocal rings, that Mn(R) has an involution does not imply that R has an anti-automorphism; an example will be given in section 11. Throughout, R is a ring, C = Cent(R), σ ∈ Aut(C), and n ∈ N is fixed. We begin by introducing some notation. For an anti-endomorphism γ : R → R, we define the standard double R-module of (R,γ) to be R endowed with the actions γ k ⊙0 r = r k and k ⊙1 r = kr. Observe that when γ is invertible, this yields a double R-progenerator of same type as γ. We now recall the following theorem, which follows from Th. 7.8 in [10] and the comment following the statement. To make the exposition more self-contained, we present here an ad-hoc proof which is slightly simpler than the proof given in [10], but not essentially different.

Theorem 7.1. Assume that up to isomorphism M = RR is the only right R-module satisfying M n =∼ Rn (e.g. when R is semilocal). Then for every anti-automorphism α : Mn(R) → Mn(R), there exists an anti-automorphism γ : R → R such that Kα is isomorphic to the standard double R-module of (R,γ). Furthermore, there is an inner automorphism ϕ ∈ Inn(Mn(R)) such that ϕ ◦ α = Tnγ where Tnγ is defined Tnγ γ by (rij ) = (rji). n n n Proof. Identify Mn(R) with EndR(R ), and let bα : R × R → K := Kα be the bilinear form induced by α. Since bα is regular, we have an isomorphism r n n [1] ∼ [1] n ∼ [1] Adbα : R → (R ) = (R ) . By assumption, this means that RR = R . [1] Observe that R = HomR(R,K0) =∼ K1 via f 7→ f(1). Thus, we may identify [1] RR =∼ R with K1, so now k ⊙1 r = kr for all k, r ∈ K = R. Define γ : R → R γ γ γ by r =1R ⊙0 r, and note that k ⊙0 r =1 ⊙0 r ⊙1 k = r ⊙1 k = r k. This easily implies that γ is an anti-endomorphism of R, and that K is just the standard double R-module of (R,γ). Since K is a double R-progenerator, we must have op op R =∼ End(K1) where the isomorphism is given by r 7→ [k 7→ k ⊙0 r]. As γ k ⊙0 r = r k, this forces γ to be an anti-automorphism. Finally, we refer the reader to the proof of [10, Th. 7.8(iv)] for the existence of ϕ.  14 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES

As a corollary, we get the following theorem, which can be regarded as a variation of Propositions 6.2.

Theorem 7.2. Suppose that up to isomorphism M = RR is the only right R-module n n satisfying M =∼ R . If Mn(R) has an involution (resp. anti-automorphism) of type σ, then M2(R) (resp. R) has an involution (resp. anti-automorphism) of type σ.

Proof. The case where Mn(R) has an anti-automorphism follows from Theorem 7.1. Let α be an involution of Mn(R). By Theorem 3.3 and Lemma 4.1, Kα is a double R-progenerator admitting an involution. Thus, by the proof of Theorem 4.5, applied [1] with P = RR, EndR(R ⊕ R ) has an involution of type σ. As in the proof of [1] [1] 2 Theorem 7.1, R =∼ (Kα)1 =∼ RR, so EndR(R ⊕ R ) =∼ EndR(R ) =∼ M2(R) (as C-algebras). 

We now show that if Mn(R) has an involution, then so does R, under further assumptions on R. To make the statement more general, we introduce reduced types. Set R := R/ Jac(R) and C := Cent(R) (C may be larger than the image of C = Cent(R) in R). Let S be a ring such that S := S/ Jac(S) is equipped with op a central C-algebra structure (e.g. R or Mn(R)). If P is an (S, R)-progenerator, then by Proposition 2.4, P Jac(R) = Jac(S)P and P := P/P Jac(R) is an (S, R)- progenerator. Thus, P induces an isomorphism Cent(R) → Cent(S). We may view this isomorphism as an automorphism σ : C → C which we call the reduced type of P . Specializing to the case S = Rop, we can define reduced types of (Rop, R)-progenerators and double R-progenerators. In addition, if α is an anti- automorphism of S, then the reduced type of α is the automorphism it induces on Cent(S) =∼ C. It is easy to check that all previous results mentioning types are also valid for reduced types, provided the following convention: If M is a right R-progenerator and S = EndR(M), then C-algebra structure of S is given by the isomorphism C → Cent(S) induced by the (S, R)-progenerator M := M/M Jac(R).

Theorem 7.3. Let R be a semilocal ring such that Mn(R) has an involution of type ∼ t σ ∈ Aut(C) and reduced type σ ∈ Aut(C). Write R := R/ Jac(R) = i=1 Mni (Di) where D is a division ring and n ∈ N, and identify C = Cent(R) with F i i Q i=1 i where F = Cent(D ). We view each F as a non-unital subring of C. Assume that i i i Q there is 1 ≤ ℓ ≤ t such that for all i ∈{1,...,t}\{ℓ} we have

(1) Di is not a field, or ni is even, or σ|Fi 6= idFi , and for the index ℓ we have × (2) Dℓ is not a field, or nℓ is even, or σ|Fℓ 6= idFℓ , or 2 ∈ Dℓ . Then R has an involution of type σ and reduced type σ.

n n Proof. Step 1. Note first that for an R-module M, M =∼ R implies M =∼ RR. Indeed, this is clear when R is semisimple, and we can reduce to this case by replacing M with M := M/M Jac(R), as explained in the proof of Theorem 5.4. Let α be an involution of Mn(R) as above. By Theorem 7.1, we may assume Kα is the standard double R-module of (R,γ) for some anti-automorphism γ : R → R, necessarily of type (resp. reduced type) σ (resp. σ). Furthermore, by Theorem 3.3, Kα has an involution θ. Namely, θ : R → R is a map satisfying (1) (aγ bc)θ = cγ bθa ∀a,b,c ∈ R. Suppose that θ fixes an invertible element u ∈ R×. We claim that the map β : R → R defined by rβ = u−1rγ u is an involution, clearly of the same type and reduced type as γ. Indeed, observe that for all r ∈ R, (ur)θ = rγ uθ = uu−1rγ u = urβ. Therefore, ur = (ur)θθ = (urβ)θ = urββ, which implies r = rββ. RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 15

We are thus reduced to show that there exists u ∈ R× with uθ = u. Likewise, we are also done if we can find u ∈ R× with uθ = −u; simply replace θ with −θ.

Step 2. Let J = Jac(R). Clearly J γ = J, hence γ induces an anti-automorphism on R. We claim that J θ ⊆ J. Indeed, J θ is an ideal by (1) and the bijectivity of γ, so is enough to show that 1+ J θ = (1θ + J)θ consists of left invertible elements. Indeed, 1 = (1 · 1)θθ = (1 · 1θ)θ = (1θ)γ · 1θ, hence 1θ is left invertible, and a simple variation shows that 1θ is also right invertible. Thus, 1θ + J consists of invertible elements. Now, θ sends right invertible elements to left invertible elements since sr = 1 implies rγ sθ = (sr)θ = 1θ ∈ R×. Therefore, (1θ + J)θ consists of left invertible elements, as required. Let γ, θ be the maps induced by γ,θ on R = R/J. In the following steps, we shall establish the existence of x ∈ R with x + xθ ∈ R× or x − xθ ∈ R×. This would finish the proof since we can then take u of Step 1 to be x ± xθ. It is enough to prove the existence of x for R, γ, θ and then lift it arbitrarily to R. Thus, we may henceforth assume R is semisimple.

Step 3. Let C = Cent(R). Then Cγ = C. Moreover, for all c ∈ C,

c = cθθ = ((1c)θ)θ = (cγ 1θ)θ = (1θcγ )θ = cγγ1θθ = cγγ , so γ|C is an involution. Let {e1,...,et} be the primitive idempotents of C. As R is semisimple, we may assume that eiRei = eiR = Mni (Di) (with Di, ni as in the theorem). Now, γ permutes {e1,...,et} and the orbits have size 1 or 2. Choose γ representatives {ei}i∈I for the orbits of γ such that 1 ∈ I, and let fi = ei + ei if γ ei 6= ei and fi = ei otherwise. Then {fi}i∈I are pairwise orthogonal, i∈I fi =1 γ and fi = fi for all i. −1 P Observe that if f ∈ R satisfies f γ = f, then (fRf)θ = f γ Rθf γ = fRf. In particular, θ takes fiRfi = fiR into itself for all i. As R = fiR, it is enough to ′ θ ′ ′θ × find xi, xi ∈ fiR such that xi + xi , xi − xi ∈ (fiR) for all i 6= ℓ, and at least one θ ′ ′θ Q θ × of xℓ + xℓ , xℓ − xℓ is a unit of fℓR. (Indeed, take x = i∈I xi if xℓ + xℓ ∈ (f1R) and take x = x′ if x′ − x′θ ∈ (f R)×.) We now split into cases. i∈I i ℓ ℓ 1 P P γ γ γ Step 4. Suppose fi = ei + ei with ei 6= ei . Then fiR = eiR × ei R. Observe θ θ γ θ γ θ γ that ei = (eiei) = ei ei ∈ ei R. We claim that ei is invertible in ei R. Indeed, we have

γ γθθ γ γ θθ γγ γθ θ γθγ γγθ γθγ θ ei = ei = (ei ei ) = (ei ei ) = ei ei = ei ei , − − − − − − − − − − − − γ γ 1θθ γ 1 γ 1 θθ γ 1θ γ 1γ 1 θ γ 1γ 1θ γ 1θγ 1 θ γ 1θγ 1 ei = ei = (ei ei ) = (ei ei ) = ei ei = ei ei ,

θ γ γγ hence ei is left and right invertible in ei R (we used (1) and the fact that ei = ei θ θ repeatedly). Since ei is invertible in eiR, it follows that ei + ei and ei − ei are γ ′ invertible in fiR = eiR × ei R. We may therefore take xi = xi = ei.

Step 5. Assume that fi = ei, and let S := fiR = eiR = Mni (Di). By Theorem 7.1, there is an anti-automorphism β : D → D and v ∈ S× such that i i ′ −1 γ Tni β ′ ′ θ v s v = s for all s ∈ S. Let γ = Tni β and define θ : S → S by s = −1 ′ ′ ′ ′ ′ ′ v−1sθvγ . It is straightforward to check that (aγ bc)θ = cγ bθ a and aθ θ = a for ′ all a,b,c ∈ S. Furthermore, for all x ∈ S, x ± xθ = v−1(vx + (vx)θ). Therefore, without loss of generality, we may replace γ, θ with γ′, θ′.

Let {ejk} be the standard Di-basis of S = Mni (Di). Then γ = Tni β implies −1 γ θ γ θ γ ejk = ekj . Therefore, (ejj Sekk) = ekkS ejj = ekkSejj , namely, θ takes ejkDi to ′ ni/2 ekj Di. Now, if ni is even, we can take xi = xi = j=1 e(2j)(2j−1). Indeed, both P 16 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES

θ ′ ′θ xi + xi and xi − xi have matrices of the form 0 ∗ ∗ 0 0 ∗ ∗ 0  .  .. with all the ∗-s being nonzero, so they are invertible. Next, embed D := Di in S by sending d ∈ D to the diagonal matrix with diagonal entries (d,...,d). Since γ = Tni β, the anti-automorphism β of D coincides with γ|D. We claim that θ sends D into itself. Indeed, as a subring of S, D is the centralizer of {ejk}, and for all d ∈ D and 1 ≤ j, k ≤ ni, we have − θ γ θ θ θ θ γ 1 θ ejkd = ekj d = (dekj ) = (ekj d) = d ekj = d ejk . ′ θ ′ ′θ × It is therefore enough to find xi, xi ∈ D with xi + xi , xi − xi ∈ D . Since D is a ′ division ring, the non-existence of xi or xi forces θ = ± idD, which in turn implies θ θ γ θ γ β = γ|D = idD, because then ±d = d = (1d) = d 1 = ±d . This is impossible in case Di is not a field or σ (which is the restriction of γ to Cent(R)) does not fix Fi = Cent(S) = Cent(D) pointwise. In case i 6= ℓ, either ni is even, or Di is not a field, or σ does not fix Fi element- ′ wise, so the existence of both xi and xi is guaranteed by the previous paragraphs. In case i = ℓ, ni is odd, Di is a field, and σ fixes Fj element-wise, we must have char Di 6= 2. Returning to the setting of the previous paragraph, this means that for any 0 6= y ∈ D, at least one of y +yθ, y −yθ is nonzero (for their sum is 2y 6= 0). ′ Thus, either x1 or x1 exists, as required.  Corollary 7.4. Let R be a semiperfect ring having an involution of type σ and ∼ t reduced type σ. Write R := R/ Jac(R) = i=1 Mni (Di) as in Theorem 7.3 and assume that there is 1 ≤ ℓ ≤ t such that for all i ∈{1,...,t}\{ℓ} we have ′ Q (1 ) Di is not a field, or σ|Fi 6= idFi , and for the index ℓ we have ′ × (2 ) Dℓ is not a field, or σ|Fℓ 6= idFℓ , or 2 ∈ Dℓ . Then the basic ring of R has an involution of type σ and reduced type σ.

Proof. Let S be the basic ring of R. By Proposition 6.2, M2(S) has an involution of type σ and reduced type σ. Using the description S given in section 6 and ∼ t Proposition 2.4, it is easy to see that S is the basic ring of R. That is, S = i=1 Di. Therefore, S satisfies the conditions of Theorem 7.3, and hence S has an involution of type σ and reduced type σ. Q  Corollary 7.5. Let R be a finite-dimensional algebra over a field F . Suppose Mn(R) admits an involution α sending F = F · 1R into itself and satisfying α|F 6= idF . Then R has an involution of the same type and reduced type as α. The same applies to the basic F -algebra that is Morita equivalent to R. Proof. Let a ∈ F be an element with aα − a 6= 0. Write C = Cent(R/ Jac(R)) = F1 ×···× Fn with each Fi a field, and let σ ∈ Aut(C) be the reduced type of α. Pick some 0 6= x ∈ Fi and view it as an element of C. If both σ(x) = x, and σ(ax) = ax, then ax = σ(ax) = σ(a)σ(x) = aαx, which implies (a − aα)x = 0, a contradiction. Thus, σ does not fix any of the fields Fi element-wise, so we are done by Theorems 7.3 and Corollary 7.4.  Remark 7.6. In general, if R, θ, γ are as in the proof of Theorem 7.3, then there may not exist u ∈ R× with u = uθ or u = −uθ. For example, take R = F ×F with F θ a field of characteristic not 2, let α = idR, and let θ be defined by (x, y) = (x, −y). This implies that the proof of Theorem 7.3 cannot be applied to arbitrary semilocal RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 17 rings. In fact, we believe that there exist semilocal rings R without involution such that M2(R) has an involution.

Remark 7.7. Given R, the datum of θ and γ satisfying (1) is equivalent to the datum of an anti-structure, a notion introduced by Wall [22, pp. 244] to define quadratic forms over rings. Recall that an anti-structure on R consists of an anti- automorphism γ : R → R and an element v ∈ R× such that vγ = v−1 and rγγ = vrv−1 for all r ∈ R. Fixing γ, there is a one-to-one correspondence between −1 the possible θ-s and v-s given by v = θ(1) and rθ = rγ v = vrγ . Therefore, the proof of Theorem 7.2 shows that if R has an anti-structure, then M2(R) has an involution. Unfolding the construction, this involution is given by

a b dγ bθ dγ bγv 7→ −1 = −1 c d v−1cθv−1 aγ v−1cγ aγ      

This trick was noted by several authors in the past (see [19, pp. 532], for instance). The proof of theorem 4.5 can be viewed as a generalization of it. This means that the question of whether there is a semilocal ring R without involution such that M2(R) has an involution is equivalent to whether there is a semilocal ring without an involution admitting an anti-structure. In fact, we were unable to find in the literature an example of a general ring with these properties.

8. Counterexamples We now demonstrate that there are “nice” rings which are not Morita equivalent to rings with an involution (of any kind), but still admit an anti-automorphism. In the first example, that automorphism fixes the center element-wise.

Example 8.1. Recall that a poset consists of a finite set I equipped with a partial order which we denote by ≤. For a field F and a poset I, the incidence algebra F (I) is defined to be the subalgebra of the I-indexed matrices over F spanned as an F -vector space by the matrix units {eij | i, j ∈ I, i ≤ j}. If I is not the disjoint union of two non-comparable subsets, then Cent(F (I)) = F . In this case, we say I is connected. The algebra F (I) is well-known to be basic. Let R be a ring that is Morita equivalent to F (I). The poset I can be recov- ered from R, up to isomorphism, as follows: Let E denote the set of primitive × idempotents in R. Then R acts by conjugation on E. Define IR to be the set of equivalence classes in E, and for i, j ∈ IR, let i ≤ j if and only if eRf 6= 0 for some (and hence any) e ∈ i and f ∈ j. To see that IR is indeed isomorphic to I, observe that there is an isomorphism between I and the isomorphism classes of in- decomposable projective R-modules via i 7→ Pi := eR (e ∈ i), and i ≤ j if and only if HomR(Pj , Pi) 6= 0. This shows that IR can be determined from Mod-R, hence ∼ IR = IF (I). Finally, IF (I) is easily seen to be isomorphic to I (via sending the class of eii ∈ F (I) to i). The construction of IR implies that any anti-automorphism (resp. involution) of R induces an anti-automorphism (resp. involution) on I =∼ IR. Suppose now that I is connected, has an anti-automorphism, but admits no involution. Then the previous paragraphs imply that Cent(F (I)) = F , and any ring that is Morita equivalent to F (I) does not have an involution. On the other hand, the anti-automorphism of I gives rise to an anti-automorphism of F (I) of type idF . An example of such a poset was given in [21] by Scharlau (for other 18 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES purposes); I is the 12-element poset whose Hasse diagram is: • / • ⑧? O _❅❅ ⑧⑧ ❅❅ ⑧⑧ ❅❅ ⑧⑧ ❅ • • ••o O

• / •• • ❅❅ ⑧ ❅❅ ⑧⑧ ❅❅ ⑧⑧ ❅ ⑧⑧ ••o   (Using Scharlau’s words, it is “the simplest example I could find”.) An anti- automorphism of I is given by rotating the diagram ninety degrees clockwise. Involutions and automorphisms of incidence algebras are well-understood in gen- eral; we refer the reader to the survey [6] and the references therein. Example 8.2. Various f.d. division algebras admitting an anti-automorphism but no involution were constructed in [16]. By Proposition 6.3, none of these algebras is Morita equivalent to a ring with involution. Note that by Albert’s Theorem ([1, Th. 10.19]), anti-automorphisms of such division algebras cannot fix the center pointwise.

9. Azumaya Algebras The rest of this paper concerns Saltman’s Theorem about Azumaya algebras with involution (see section 1) and the construction of Azumaya algebras without involution. As preparation, we now briefly recall Azumaya algebras and several facts about them to be used later. We refer the reader to [20] and [8] for an extensive discussion and proofs. Throughout, C is a commutative ring and, unless specified otherwise, all tensor products are taken over C. 9.1. Azumaya Algebras. A C-algebra A is called Azumaya if A is a progenerator op as a C-module and the standard map Ψ : A ⊗ A → EndC (A) given by Ψ(a ⊗ bop)(x) = axb is an isomorphism. When C is a field, being Azumaya is equivalent to being simple and central, so Azumaya algebras are a generalization of central simple algebras. Let A, B be Azumaya C-algebras. The following facts are well-known:

(1) Cent(A)= C and CentA⊗B(C ⊗ B)= A ⊗ C. (2) A ⊗ B and Aop are Azumaya C-algebras. (3) If ψ : C → C′ is a commutative , then A ⊗ C′ is an Azumaya C′-algebra (C′ is viewed as a C-algebra via ψ). (4) For every C-progenerator P , EndC (P ) is an Azumaya C-algebra. ′ (5) If B is a subalgebra of A, then B := CentA(B) is Azumaya (over C), ′ ′ ′ ′ B = CentA(B ) and B ⊗ B =∼ A via b ⊗ b 7→ bb .

For every M ∈ proj-C, we define rank(M) = rankC (M) to be the function

Spec(C) → Z sending a prime ideal P to dimkP (M ⊗ kP ) where kP is the fraction field of C/P . We write rank(M) = n to denote that rank(M)(P ) = n for all P ∈ Spec(C). For example, when C is a field, rank(M) = dimC (M). More generally, rank(M) is constant when Spec(C) is connected (as a topological space), or equivalently, when C does not decompose as product of two nonzero rings. In (M) (M) (M) fact, we can always write C = C1 ×···× Ct such that rank(M ⊗ Ci ) is constant. (This follows from the well-known facts that rank(M) : Spec(C) → Z is continuous, and Spec(C) is a compact topological space.) Lastly, observe that RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 19 if ψ : C → C′ is a commutative ring homomorphism and P ′ ∈ Spec(C′), then ′ ′ −1 ′ rankC′ (M ⊗ C )(P ) = rankC (M)(P ) where P = ψ (P ). This allows us to extend scalars when computing ranks. Proposition 9.1. Let A and B be Azumaya C-algebras and let ϕ : A → B be a homorphism of C-algebras. Then ϕ is injective. If moreover rank(A) = rank(B), then ϕ is an isomorphism.

Two Azumaya C-algebras A, B are said to be Brauer equivalent, denoted A ∼Br B, if there are C-progenerators P,Q such that

A ⊗ EndC (P ) =∼ B ⊗ EndC (Q) as C-algebras. The Brauer class of A, denoted [A], is the collection of Azumaya algebras which are Brauer equivalent to A. The Brauer group of C, denoted Br(C), is the set of all Brauer equivalence classes endowed with the group operation [A] ⊗ [B] = [A ⊗ B]. The unit element of Br(C) is the class of C, namely, the class of Azumaya algebras which are isomorphic to EndC (P ) for some progenerator P . The inverse of [A] is [Aop]. The following theorem, which is essentially due to Bass, presents an alternative definition of the Brauer equivalence.

Theorem 9.2 (Bass). Let A, B be two Azumaya C-algebras. Then A ∼Br B ⇐⇒ A ∼Mor/C B, i.e. there exists an (A, B)-progenerator P such that cp = pc for all p ∈ P and c ∈ C. Proof. See [3, Cor. 17.2]. The assumption that Spec(C) is noetherian in [3] can be ignored by [20, Th. 2.3]; see also [19, §1].  The exponent of an Azumaya algebra A is its order in Br(C). Theorem 9.3. Let A be an Azumaya C-algebra. Then rank(A) is a square. Fur- 2 ⊗n thermore, if rank(A) divides n , then A ∼Br C. We will also need the following proposition. Proposition 9.4. Let A, B be Azumaya C-algebras and let K be an (A, B)- bimodule satisfying ck = kc for all c ∈ C. Then K is an (A, B)-progenerator if and only if K is a (C, Aop ⊗ B)-progenerator with Aop ⊗ B acting via k(aop ⊗ b)= akb. op In particular, in this case, EndC (C K) =∼ A ⊗ B. op Proof. The “only if” part easily follows from CentAop⊗B(C ⊗ B)= A ⊗ C, so we turn to prove the “if” part. Let D = A ⊗ Aop and consider A as a right D-module op via a(x ⊗ y ) = yax. Since C A is a progenerator and D =∼ End(C A) (because A is Azumaya), A is a (C,D)-progenerator. By Proposition 2.3, K ⊗ Aop is an op op op op (A ⊗ A ,B ⊗ A )-progenerator and hence C A ⊗D (K ⊗ A ) is a (C,B ⊗ A )- op progenerator. It is straightforward to check that C A ⊗D (K ⊗ A ) =∼ K as (C,B ⊗ op op A )-bimodules via x ⊗D (k ⊗ y ) 7→ yxk and k 7→ 1 ⊗D (k ⊗ 1) in the other direction, so we are done.  9.2. Galois Extensions. Let G be a finite group of ring-automorphisms of C and G 2 × 4 let C0 = C be the ring of elements fixed by G. For every 2-cocycle f ∈ Z (G, C ), let ∆(C/C0,G,f) = σ∈G Cuσ where {uσ}σ∈G are formal variables. We make ∆(C/C0,G,f) into a ring by linearly extending the following relations L uσuτ = f(σ, τ)uστ , uσc = σ(c)uσ, ∀σ, τ ∈ G, c ∈ C.

4 Recall that a 2-cocycle of G taking values in C× is a function f : G × G → C× satisfying σ(f(τ, η))f(σ, τη) = f(σ, τ)f(στ,η) and f(1, σ) = f(σ, 1) = 1 for all σ, τ, η ∈ G. The 2-cocycles with the operation of point-wise multiplication form an abelian group denoted Z2(G, C×). 20 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES

This makes ∆(C/C0,G,f) into a C0-algebra whose unity is u1. We say that C/C0 is a Galois extension with Galois group G if ∆(C/C0, G, 1) is an Azumaya C0-algebra (1 denotes the trivial 2-cocycle f(σ, τ) ≡ 1). In this case: ∼ (1) ∆(C/C0, G, 1) = EndC0 (C) via σ cσuσ 7→ [x 7→ σ cσσ(x)] ∈ EndC0 (C). 2 × (2) The C0-algebra ∆(C/C0,G,f) is Azumaya for all f ∈ Z (G, C ). P P (3) rankC0 (C)= |G|.

(4) For every commutative C0-algebra D, the extension C ⊗C0 D/C0 ⊗C0 D is

Galois with Galois group {g ⊗C0 idD | g ∈ G}. See [20, §6] for proofs. Azumaya algebras of the form ∆(C/C0,G,f) are called crossed products.

Proposition 9.5. Let C/C0 be a Galois extension with Galois group G. Assume M is a C-module endowed with a G-action such that σ(mc) = σ(m)σ(c) for all G ∼ σ ∈ G, m ∈ M, c ∈ C. Then M ⊗C0 C = M via m ⊗ c 7→ mc, where M G := {m ∈ M : σ(m)= m ∀σ ∈ G} . G G In particular, rankC0 (M) = rankC0 (C)rankC0 (M )= |G| · rankC0 (M ). Further- more, if M is an Azumaya C-algebra and G acts via ring automorphisms on M, G then M is an Azumaya C0-algebra. Proof. See [20, Prps. 6.10 & 6.11]. 

9.3. Corestriction. Let C/C0 be a Galois extension with Galois group G. For every Azumaya C-algebra A and σ ∈ G, define Aσ to be the C-algebra obtained from A by viewing A as a C-algebra via σ : C → A. The algebra Aσ is also Azumaya, but it need not be Brauer-equivalent to A. Observe that the algebra σ − B := σ∈G A admits a G-action given by τ( σ aσ)= σ aτ 1σ and that action satisfies σ(bc)= σ(b)σ(c) for all σ ∈ G, b ∈ B, c ∈ C. The corestriction of A (with N N N respect to C/C0) is defined to be G G σ CorC/C0 (A)= B = A .  σO∈G 

By Proposition 9.5, CorC/C0 (A) is an Azumaya C0-algebra. Moreover, it can be shown that the corestriction induces a group homomorphism

CorC/C0 : Br(C) → Br(C0) given by CorC/C0 ([A]) = [CorC/C0 (A)]. We also note that the corestriction can be defined for separable extensions C/C0; see [20, Ch. 8] for further details.

10. Saltman’s Theorem In this section, we show how to recover the Knus-Parimala-Srinivas proof of Saltman’s Theorem ([13, §4]) as an application of Theorem 4.5. In order to keep the exposition as self-contained and fluent as possible, we will first give a proof of Saltman’s Theorem using Theorem 4.5, and then explain how this proof relates to the proof in [13]. 10.1. Saltman’s Theorem. Let C be a commutative ring and let R be a C- algebra. Recall that a Goldman element of R/C is an element g ∈ R ⊗C R such that g2 = 1 and g(r ⊗ s) = (s ⊗ r)g for all r, s ∈ R. Proposition 10.1. All Azumaya C-algebras have a Goldman element Proof. See [14, p. 112] or [20, Pr. 5.1], for instance.  Theorem 10.2 (Saltman). Let C be a commutative ring and let A be an Azumaya C-algebra. Then: RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 21

(i) A is Brauer equivalent to an Azumaya algebra B with an involution of the op first kind if and only if A ∼Br A . (ii) Let C/C0 be a Galois extension with Galois group {1, σ} (σ 6= 1). Then A is Brauer equivalent to an Azumaya algebra B with an involution whose

restriction to C is σ if and only if CorC/C0 (B) ∼Br C0. op Proof. (i) By Theorem 9.2, A ∼Br A if and only if there exists a double A- op progenerator of type idC (i.e. an (A , A)-progenerator of type idC ). Likewise, by Theorems 4.5 and 9.2, A is Brauer equivalent to a C-algebra B with involution of the first kind if and only if there exist a double A-progenerator of type idC with involution. Therefore, it is enough to show that every double A-progenerator K of type idC has an involution. Indeed, consider K as an A ⊗C A-module via ′ ′ ′ k · (a ⊗C a )= k ⊙0 a ⊙1 a (a,a ∈ A) and let g be a Goldman element of A. It is easy to check that k 7→ kg is an involution of K. σ G (ii) Throughout, D = A ⊗C A and E = D = CorC/C0 (D). Recall that σ extends to an automorphism of D, also denoted σ, given by σ(a ⊗ a′)= a′ ⊗ a. By

Theorems 4.5 and 9.2, it is enough to prove that CorC/C0 (A) ∼Br C0 if and only if then there exists a double A-progenerator of type σ admitting an involution. Assume K is a double A-progenerator of type σ with involution θ. Then K can op σ be considered as an (A , A )-progenerator of type idC , hence by Proposition 9.4, σ End(C K) =∼ A ⊗ A = D. Extend the left action of C on K to a left action of ∆ := ∆(C/C0, G, 1) on K by letting uσ act as θ. Then it is easy to see that ∼ G op End(∆K) = D = E. Viewing ∆ and E as subrings of End(C0 K), this means E is op the centralizer of ∆ in End(C0 K). As all these algebras are Azumaya over C0, we op ∼ op have ∆ ⊗C0 E = End(C0 K). InBr(C0) this reads as [C0]⊗[E] = [∆ ]⊗[E] = [C0], so CorC/C0 (A)= E ∼Br C0.

Conversely, assume E = CorC/C0 (A) ∼Br C0. Then there exists a (C0, E)- progenerator Q of type idC0 . Let K = Q ⊗C0 C. Then by Proposition 2.2, K is a

(C0⊗C0 C, E⊗C0 C)-progenerator. Identify E⊗C0 C with D via (e⊗c) 7→ ec. Then K is a (C,D)-progenerator, hence by Proposition 9.4, K is an (Aop, Aσ)-progenerator via aop · k · a′ := k(a ⊗ a′). We may thus view K as a double A-progenerator of θ type σ. Let θ = σ ⊗C0 idQ : K → K (so (c ⊗ q) = σ(c) ⊗ q). We claim that θ is 2 ′ an involution of K. Indeed, θ = idK and for all c,c ∈ C, e ∈ E, q ∈ Q, we have ((c ⊗ q)(c′e))θ = (cc′ ⊗ qe)θ = σ(c)σ(c′) ⊗ qe = (σ(c) ⊗ q)(σ(c′)e) = (c ⊗ q)θ(σ(c′e)) (e = σ(e) since e ∈ E = DG). This implies that (kx)θ = kθσ(x) for all k ∈ K, x ∈ D. Putting x =1 ⊗ a and x = a ⊗ 1 yields that θ is an involution.  Remark 10.3. (i) The proof of Theorem 10.2 also shows that a C-algebra R with a Goldman element is Morita equivalent to its opposite over C if and only if it is Morita equivalent as a C-algebra to an algebra with an involution of the first kind. We could not find a non-Azumaya algebra admitting a Goldman element, though. (ii) A slightly different proof of the “only if” part of Theorem 10.2(ii) that does not use double progenerators appears in [19, pp. 531]. In order to explain the connection of the previous proof with the Knus-Parimala- Srinivas proof of Saltman’s Theorem ([13, §4]), let us backtrack our proof to see [1] what is the algebra B. According to Theorem 4.5, B = EndA(P ⊕ P ) with P an arbitrary A-progenerator, and the involution of B is induced by the bilinear form bP constructed in the proof of Theorem 4.5. The functor [1] is computed using a double A-progenerator K with involution which is obtained in the proof of [1] Theorem 10.2. Let us choose P = AA. Then B =∼ EndA(A ⊕ K1) (since A = HomA(AA,K0) =∼ K1 via f 7→ f(1A)). This algebra with involution is essentially the algebra with involution constructed in [13, §4]. However, in [13], the involution of B is constructed directly, while here we have obtained it from a general bilinear 22 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES form on the right A-module A⊕K1, and hence suppressed some of the computations of [13]. We also note that many proofs of special cases of Saltman’s Theorem in the literature, [13] and [19] in particular, involve the construction of an involutary map θ taking some (Aop, A)-bimodule or an equivalent object K into itself and satisfying (aopkb)θ = bopkθa (a,b ∈ A, k ∈ K). For example, consider the map α in [19, pp. 532–3, 537], the maps ψ and σP in [13, pp. 71–2] and the map u of [12, pp. 196]. This hints that these proofs can be effectively phrased using general bilinear forms. 10.2. Rank Bounds. We proceed by showing that the algebra B in Saltman’s Theorem can be taken to have rank(B) = 4 rank(A) in part (i) and rank(B) = 4 min{rank(A), rank(Aσ)} in part (ii). This follows from the following technical lemmas. Lemma 10.4. Let A be an Azumaya C-algebra and let M,N ∈ proj-A. Then rank(M) · rank(N) rank(Hom (M,N)) = . A rank(A)

Proof. Let P ∈ Spec(C) and let F be an algebraic closure of kP , the fraction field of C/P . Set AF = A⊗C F , MF = M ⊗C F and NF = N ⊗C F . By Proposition 2.2, ∼ ∼ we have HomAF (MF ,NF ) = HomA(M,N) ⊗C F = (HomA(M,N) ⊗C kP ) ⊗kP F . Therefore, it is enough to verify the lemma in case C is an algebraically closed field, i.e. when C = F , in which case rankC (−) and dimF (−) coincide. In this case, we must have A =∼ Mn(F ) (as F -algebras) and M =∼ Ms×n(F ), N =∼ Mt×n(F ) (as A-modules) for suitable s,t,n ∈ N ∪ {0}. But then, dimF HomA(M,N) = dimF Mt×s(F )= ts = dimF (N) · dimF (M)/ dimF (A), as required.  Lemma 10.5. Let A be an Azumaya C-algebra and let K be a double A-progenerator of type σ ∈ Aut(C). Then

σ σ−1 rank(K0)= rank(A) rank(A ) and rank(K1)= rank(A) rank(A ) q Proof. ConsiderpK as an (Aop, A)-bimodule as explained in section 4. Since A is σ op of type σ, EndA(K0) = End(KA) =∼ (A ) as C-algebras, hence rank(EndA(K0)) = σ 2 rank(A ). However, by Lemma 10.4, we have rank(EndA(K0)) = rank(K0) / rank(A). Comparing both expressions yields the formula for rank(K0). The formula for rank(K1) is shown in the same manner.  Proposition 10.6. In Theorem 10.2(i) (resp. Theorem 10.2(ii)), the algebra B can be chosen such that rank(B) = 4 rank(A) (resp. rank(B) = 4 min{rank(A), rank(Aσ)}).

Proof. Let σ be idC in the case of Theorem 10.2(i) and a nontrivial Galois automor- phism of C/C0 in the case of Theorem 10.2(ii). In the comment after Remark 10.3 we saw that B can taken to be EndA(A ⊕ K1) with K a double A-progenerator of type σ. In this case, by Lemmas 10.5 and 10.4, we have 2 σ 2 rank(A)+ rank(A) rank(A ) rank(Aσ) rank(B)= = 1+ rank(A) .  prank(A)  s rank(A) ! In the case of Theorem 10.2(i), this means rank(B) = 4 rank(A), so we are done. For the case of Theorem 10.2(ii), it is enough to show that A is Brauer equivalent to an Azumaya algebra A′ of rank min{rank(A), rank(Aσ)} (for then we can replace A with A′ and get the desired rank). Recall from section 9 that we can decompose C as a product C = C1 ×···× Ct σ such that rank(A ⊗C Ci) and rank(A ⊗C Ci) are constant for all i. Let ei ∈ C RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 23 be the idempotent satisfying Ci = eiC. We identify Spec(C) with Spec(Ci) in the standard way. It is easy to see that every C-algebra R factors as a product S i eiR and R/C is Azumaya if and only if eiR/eiC is Azumaya for all i. Define ′ ′ ′ σ ′ σ op A = i Ai where Ai = eiA if rankCi (eiA) < rankCi (ei(A )) and Ai = ei(A ) Q ′ ′ otherwise. Then A ∼Br A (since eiA ∼Br eiA for all i), and, by definition, Q ′ σ rankC (A ) = min{rankC (A), rankC (A )}, as required.  10.3. The Semilocal Case. In [19, §4], Saltman also showed that when C is semilocal and connected, one can take B = A in Theorem 10.2 (Albert’s Theorems follow as the special case where C is a field). Let us recover the proof using the language and the tools of the previous sections: Since C is connected, A has constant rank, say rank(A) = n2. Suppose the assumptions of parts (i) or (ii) of Theorem 10.2 are satisfied, and set σ = idC for part (i). By the proof of Theorem 10.2, there is a double A-module K of type σ admitting an involution θ. We claim that K is the standard double A-module of some anti-automorphism γ : A → A (see section 7). Indeed, by assumption, 2 σ rank(A) = n = rank(A ), hence by Lemma 10.5, rank(K1) = rank(A). We now claim that for P,Q ∈ proj-A, rank(P ) = rank(Q) implies P =∼ Q. By tensoring with C/ Jac(C), we may assume C is a finite product of fields and A is a finite product of central simple algebras, in which case the claim is routine. Therefore, K1 =∼ AA, and arguing as in the proof of Theorem 7.1 yields the desired γ. We can now apply the proof of Theorem 7.3 with R = A (which is semilocal) to assert that A has an involution of type σ. However, we need to verify the assumptions of the theorem, so let ni, Di and Fi be defined as in Theorem 7.3 for 2 2 R = A. Since rank(A) = n , we have dimFi Mni (Di) = n for all i. Thus, when n is even, for all i, either ni is even or Di is a division ring, as required. When n is odd and σ 6= idC, set A = A/ Jac(A), C = C/ Jac(C) and C0 = C0/ Jac(C0). It is a standard fact that A/C is Azumaya (so C = Cent(A)) and C/C0 is Galois with Galois group {σ, id}, where σ is the map induced by σ on C. Noting that C = F ×···× F and rank (C) = 2, we see that σ cannot fix any F pointwise, so 1 t C0 i again the assumptions of Theorem 7.3 hold. Finally, in case n is odd and σ = idC , ⊗2 ⊗n we have A ∼Br C (by assumption) and A ∼Br C (by Theorem 9.3, since 2 rank(A)= n ), which implies A ∼Br C, i.e. A =∼ EndC (P ) for some C-progenerator P , necessarily of rank n. By what we have shown above, this implies P =∼ Cn, so A =∼ EndC(P ) =∼ Mn(C) and A has the transpose involution. We note that specializing in Theorem 7.3 to the case R is an Azumaya C- algebra and C is connected saves many technicalities in the last steps of the proof. In particular, one can use the Skolem-Noether Theorem rather than basing on Theorem 7.1; cf. the argument given by Saltman in [19, pp. 538]. Remark 10.7. Assume C is semilocal or Q-finite and let τ be any automorphism of C. Then Corollary 5.6 (together with Theorem 9.2) implies that A is Brauer equivalent to an Azumaya algebra with an anti-automorphism of type τ if and only op τ if A ∼Br A . The semilocal part of this claim is somewhat standard since it is op τ op τ well-known that A ∼Br A implies A =∼ A in this case (e.g. see the previous paragraphs). However, the Q-finite case seems to be unknown. When τ is a Galois automorphism of order 2 (i.e. C/C{τ} is Galois with Galois group {1, τ}), the op τ τ ∼ condition A ∼Br A is equivalent to A ⊗C A = ResC/C0 (CorC/C0 (A)) ∼ C. In this special case, it could be that the claim is true without any assumption on C.

11. Azumaya Algebras of Exponent 2 Without Involution

Let C be a commutative ring and let A be an Azumaya C-algebra with A ∼Br op A . Then Saltman’s Theorem asserts that there is some algebra B ∼Br A with 24 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES involution of the first kind. However, it can be very tricky to construct examples of op algebras A with A ∼Br A that do not have an involution already. For example, Saltman have shown in [19, §4] that this is impossible if C is semilocal or rank(A)= 4. In particular, A cannot be a tensor product of quaternion algebras. (The proof of the semilocal case was recovered in subsection 10.3. In case rank(A) = 4, the map a 7→ Tr(a) − a is an involution on A, where Tr(a) denotes the reduced trace of a.) Saltman has pointed out in [19, §4] how to produce examples of Azumaya algebras without involution that are trivial in Br(C). However, we have not seen in the literature examples of non-trivial such algebras. Of course, one can always take the product of a non-trivial quaternion algebra with a trivial Azumaya algebra without involution, but this results in C being non-connected (i.e. C is a product of two rings). One is therefore interested in examples in which C is connected. More generally, one can ask whether the bound rank(B) ≤ 4 rank(A) of Proposition 10.6 is tight. In this section, we shall use the results of the section 4 to explain how to detect whether an Azumaya algebra admits an anti-automorphisms. This will be used to construct an explicit example of a non-trivial Azumaya algebra over a Dedekind domain C that does not have an involution. However, this example does not show that the bound rank(B) ≤ 4 rank(A) is optimal, and it also has zero divisors. (Even worse, we shall see that for C as in the example, and also for many other low-dimensional domains, we can always choose B to be without zero divisors and to satisfy rank(B) ≤ rank(A).) We thank Asher Auel for pointing us to this problem. Throughout, C is a commutative ring and σ ∈ Aut(C). Unless specified other- wise, all tensor products are taken over C. Recall that the Picard group, Pic(C), consists of the isomorphism classes of rank-1 projective C-modules, multiplication being tensor product (over C). Equivalently, Pic(C) consists of isomorphism classes of (C, C)-progenerators of type idC . In case C is a Dedekind domain (e.g. a number ring), Pic(C) is just the ideal class group of C (e.g. see [17]).

op σ Theorem 11.1. Let A be an Azumaya C-algebra such that A ∼Br A , let K op σ be any (A , A )-progenerator of type idC (cf. Theorem 9.2), which we also view as a double A-progenerator (of type σ), let P be a right A-progenerator and set [1] P = HomA(P,K0). Then the following conditions are equivalent:

(a) B := EndA(P ) has an anti-automorphism of type σ. (b) There is [I] ∈ Pic(C) such that I ⊗ P =∼ P [1] as A-modules. [1] (c) EndA(P ) =∼ EndA(P ) as C-algebras. [1] In particular, if I ⊗ P ≇ P for all [I] ∈ Pic(C), then EndA(P ) does not have an anti-automorphism of type σ.

Proof. (a)=⇒(b): By Proposition 4.6, there exists a double A-progenerator K′ of type σ and a regular bilinear form b : P × P → K′. View K′ as a right A ⊗ Aσ- ′ ′ σ module via k(a ⊗ a )= k ⊙0 a ⊙1 a and view K as a right A ⊗ A in the standard way. By Proposition 9.4, both K and K′ are (C, A⊗Aσ ) progenerators. Thus, there ′ ′ exists a (C, C)-progenerator I such that K = I ⊗C K. As both K and K have type σ ∈ Aut(C), I is of type idC and hence [I] ∈ Pic(C). Now, by Proposition 2.1, we ∼ ′ ∼ ∼ have P = Hom(P,K0) = HomC⊗A(C ⊗P, I ⊗K0) = HomC (C, I)⊗HomA(P,K0) = I ⊗ P [1]. (b)=⇒(c): We have EndC (I)= C since I is a (C, C)-progenerator of type idC , so [1] by Proposition 2.1, EndA(P ) =∼ EndC⊗A(I ⊗ P, I ⊗ P ) =∼ EndC(I) ⊗ EndA(P ) =∼ C ⊗ EndA(P ) =∼ EndA(P ). The isomorphism is easily seen to be a C-algebra isomorphism. RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 25

[1] (c)=⇒(a): Let ψ : EndA(P ) → EndA(P ) be an isomorphism. The map [1] [1] w 7→ w : EndA(P ) → EndA(P ) is an anti-homomorphism of rings (since [1] is an additive contravariant functor). Moreover, it is an anti-isomorphism of rings since [1] : proj-A → proj-A is a duality of categories (Lemma 4.2(ii)). It is also [1] straightforward to check that for c ∈ C ⊆ EndA(P ), one has c = σ(c). Thus, α [1] α : EndA(P ) → EndA(P ) defined by w = ψ(w ) is an anti-automorphism of type σ. 

Remark 11.2. By applying Theorem 11.1 with P = AA, we can test whether A [1] itself has an involution. In this case, P is just K1 (via f 7→ f(1)), so we only need to verify that I ⊗ A ≇ K1 as A-modules for all [I] ∈ Pic(C). We shall now use Theorem 11.1 to construct an Azumaya algebra B over a Dedekind domain C such that B ⊗ B ∼Br C, B ≁Br C and B does not have a C-anti-automorphism. For a fractional ideal I of C, we shall write I⊗n for the n-th power of I as a fraction ideal, and I⊕n for the C-module I ⊕···⊕ I (n times). The ambiguous notation In shall be avoided henceforth, except for writing I−1 to denote the inverse of I. We shall make extensive usage of some well-known facts about projective modules over Dedekind domain, such as: Every f.g. projective module over a Dedekind domain C is a direct sum of fractional ideals, and for n ∼ n fractional ideals I1,...,In,J1,...,Jn, we have i=1 Ii = i=1 Ji if and only if [ Ii] = [ Ji] in Pic(C). We refer the reader to [17] for proofs. i i L L ExampleQ Q 11.3. Let C is a Dedekind domain admitting a non-trivial quaternion Azumaya C-algebra A. By [19, Th. 4.1], A has an involution of the first kind α, which gives rise to a double A-module K with K1 =∼ AA (see section 7). Viewing ⊕3 A as a C-module, we can write AC =∼ C ⊕ J with J a fractional ideal. Suppose further that there is [L] ∈ Pic(C) such that [L⊗8] is not a 16-th power in Pic(C). We claim that the Azumaya C-algebra ⊕3 B := EndC (C ⊕ L) ⊗ A, does not have an anti-automorphism of type idC, despite clearly satisfying B ⊗ B ∼Br C and B ≁Br C. We show this by applying Theorem 11.1 with P = ⊕3 (C ⊕ L) ⊗ AA. (The module P is a right C ⊗ A-module which we view as a right A-module in the standard way. We have EndA(P ) =∼ B by Proposition 2.1.) In fact, we shall show that I ⊗ P ≇ P [1] as C-modules for all [I] ∈ Pic(C). [1] ⊕3 Indeed, observe that P = HomA(P,K0) = HomC⊗A((C ⊕L)⊗AA, C⊗K0) =∼ ⊕3 ⊕3 −1 HomC (C ⊕ L, C) ⊗ HomA(AA,K0) =∼ (C ⊕ L ) ⊗ K1. Thus, we have [1] ⊕3 −1 ⊕3 −1 ⊕3 P =∼ (C ⊕ L ) ⊗ K1 =∼ (C ⊕ L ) ⊗ (C ⊕ J) =∼ C⊕9 ⊕ (L−1)⊕3 ⊕ J ⊕3 ⊕ L−1J =∼ C15 ⊕ (L−1)⊗4J ⊗4 . (as C-modules). On the other hand, for [I] ∈ Pic(C), I ⊗ P = I ⊗ (C⊕3 ⊕ L) ⊗ A =∼ I ⊗ (C⊕3 ⊕ L) ⊗ (C⊕3 ⊕ J) =∼ I⊕9 ⊕ (IJ)⊕3 ⊕ (IL)⊕3 ⊕ (ILJ) =∼ C⊕15 ⊕ I⊗16L⊗4J ⊗4 . It follows that I⊗P =∼ P [1] as C-modules if and only if [I⊗16L⊗4J ⊗4] = [(L−1)⊗4J ⊗4] in Pic(C), or equivalently, if [(I−1)⊗16] = [L⊗8]. But this is impossible by the as- sumption on L. We remark that if L is chosen such that [L⊗16] = [C], then (C⊕3 ⊕L)⊕16 =∼ C⊕64. In this case, M16(B) =∼ A ⊗ M16(C), so M16(B) has an involution of type idC . As a result, one of the algebras B, M2(B), M4(B), M8(B) does not an involution of type idC while its 2 × 2 matrix algebra does have such involution. In addition, if Aut(C) = idC (e.g. if the fraction field of C has no nontrivial automorphisms), then B has no anti-automorphisms at all. 26 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES

Explicit choices of C, A and L satisfying all previous conditions (including ⊗16 [L ] = [C] and Aut(C) = idC) are the following: Let D be the integer ring of K := Q[x | x3 + x +521 = 0], and take C = D[2−1], A = C[i, j | i2 = j2 = −1, ij = −ji],L = h113, x − 16i⊗3 .

That A is Azumaya follows from the general fact (left to the reader) that (a,b)2,C := 2 2 × 1 C[i, j | i = a, j = b, ij = −ji] is Azumaya for all a,b ∈ C , provided 2 ∈ C. The algebra A is non-trivial in Br(C) since K (and hence C) embeds in R, and A ⊗C R =∼ (−1, −1)2,R is well-known to be a division ring. The class group of D is a cyclic group of order 48 generated by [h113, x − 16i] (verified using SAGE), so when viewed as an element of Pic(D), [h113, x − 16i⊗3] has order 16 and its eighth 1 power is not a 16-th power in Pic(D). Adjoining 2 to D does not affect these facts by the following easy lemma (note that 2 is prime in D). Lemma 11.4. Let D be a Dedekind domain, let 0 6= p ∈ D be a prime element of 1 D, and set Dp = D[ p ]. Then the map [I] 7→ [IDp] : Pic(D) → Pic(Dp) (known as the restriction map) is an isomorphism of groups.

Proof. That [I] 7→ [IDp] is a group homomorphism is straightforward. It is onto since any ideal I E Dp is easily seen to satisfy (I ∩ D)Dp = I. To show injectivity, observe that D has Krull-dimension 1, hence pD is maximal. As a result, every ideal 0 6= J E D can be written as J = pnJ ′ (n ∈ N) with J ′ is coprime to p. The same applies to elements x ∈ D. Also observe that if J is coprime to p and x ∈ D, the px ∈ J implies x ∈ J. Indeed, px ∈ J ∩ pD = pJ (since J is coprime to p), hence x ∈ J. Suppose now that J E D is such that I := JDp is principal. We need to show that J is principal (i.e. that [J] is trivial in Pic(D)). Replacing J with p−nJ if needed, we may assume that J is coprime to p. Let z be a generator of I. Again, replacing z with pnz (n ∈ Z), we may assume that z ∈ J, and since J is coprime to p, we may further divide z by p unitl it is coprime to p. Let y ∈ J. Then y = z(pnw) for some n ∈ Z and w ∈ D coprime to p. This implies p−ny = zw, so n cannot be negative (since the right hand side is not divisible by p). Therefore, wpn ∈ D and y = zwpn ∈ zD. This means that J = zD, as required. 

Remark 11.5. Let C be a regular integral domain of Krull dimension at most 2, and let F be the fraction field of C. Assume further that every central simple F -algebra of exponent 2 is Brauer equivalent to a quaternion algebra. For exam- ple, this is the case when F is a global field by the Albert-Brauer-Hasse-Noether Theorem, or when F has transcendence degree at most 2 over a separably closed subfield by [7]. Then every Azumaya C-algebra A of exponent 2 admits an algebra B ∈ [A] with involution and rankC (B) ≤ rankC (A). In particular, in this case, the bound rankC(B) ≤ 4 rankC (A) of Proposition 10.6 is not tight. Indeed, by Saltman [19, Th. 4.1], it is enough to show that B can be chosen to be a quaternion algebra. ′ As A := A ⊗C F has exponent 2, there exists by assumption a quaternion central ′ ′ ′ simple F -algebra B ∼Br A . Now, by the proof of [2, Pr. 7.4], B contains an ′ Azumaya C-algebra B such that B = B ⊗C F . Finally, by [2, Th. 7.2], B ∼Br A, so we are done.

12. Questions We finish with several questions that we were unable to answer.

op Question 1. Is there a ring R such that R ∼Mor R but R is not Morita equivalent to a ring with an anti-automorphism (cf. section 1)? RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES 27

Question 2. Is there a semilocal ring R such that R does not have an involution despite M2(R) having an involution (cf. Theorem 7.3)? Question 3. Is there a non-Azumaya algebra admitting a Goldman element (cf. Remark 10.3(i))?

Question 4. Let C/C0 be a Galois extension with Galois group {1, σ}. Is it true σ that all Azumaya C-algebras A with A ∼Br A are Brauer equivalent to an Azumaya algebra with an anti-automorphism whose restriction to C is σ (cf. Remark 10.7)? Question 5. Is the bound rank(B) ≤ 4 rank(A) of Proposition 10.6 tight when the base ring is connected (cf. Remark 11.5)? Question 6. Are there Azumaya algebras of exponent 2 and without zero-divisors that do not admit an involution of the first kind? For our last question, recall that an involution α on a ring R is called hyperbolic if there exists an idempotent e ∈ R such that e + eα = 1. In all proofs of Salt- man’s Theorem that we have encountered (e.g. [19], [13], and the proof given in section 10), the constructed involution was hyperbolic. In contrast, Albert’s The- orem guarantees an involution which is non-hyperbolic (and in fact anisotropic); indeed, we can take B to be a division ring, and any involution on a division ring is non-hyperbolic. We therefore ask:

Question 7. Let A be an Azumaya C-algebra with A⊗A ∼Br C. Does there always exist an Azumaya algebra B ∈ [A] having a non-hyperbolic involution of the first kind? References

[1] A. Adrian Albert. Structure of algebras. Revised printing. American Mathematical Society Colloquium Publications, Vol. XXIV. American Mathematical Society, Providence, R.I., 1961. [2] Maurice Auslander and Oscar Goldman. Maximal orders. Trans. Amer. Math. Soc., 97:1–24, 1960. [3] H. Bass. K-theory and stable algebra. Inst. Hautes Etudes´ Sci. Publ. Math., (22):5–60, 1964. [4] Hyman Bass. Finitistic dimension and a homological generalization of semi-primary rings. Trans. Amer. Math. Soc., 95:466–488, 1960. [5] E. Bayer-Fluckiger, C. Kearton, and S. M. J. Wilson. Decomposition of modules, forms and simple knots. J. Reine Angew. Math., 375/376:167–183, 1987. [6] Rosali Brusamarello and David W. Lewis. Automorphisms and involutions on incidence al- gebras. Linear Multilinear Algebra, 59(11):1247–1267, 2011. [7] A. J. de Jong. The period-index problem for the Brauer group of an algebraic surface. Duke Math. J., 123(1):71–94, 2004. [8] Frank DeMeyer and Edward Ingraham. Separable algebras over commutative rings. Lecture Notes in Mathematics, Vol. 181. Springer-Verlag, Berlin, 1971. [9] Alberto Facchini and Dolors Herbera. Projective modules over semilocal rings. In Algebra and its applications (Athens, OH, 1999), volume 259 of Contemp. Math., pages 181–198. Amer. Math. Soc., Providence, RI, 2000. [10] Uriya A. First. General bilinear forms. Israel Journal of Mathematics, 2014. Currently avail- able via Springer Link. [11] I. N. Herstein. Rings with involution. The University of Chicago Press, Chicago, Ill.-London, 1976. Chicago Lectures in Mathematics. [12] Nathan Jacobson. Finite-dimensional division algebras over fields. Springer-Verlag, Berlin, 1996. [13] M.-A. Knus, R. Parimala, and V. Srinivas. Azumaya algebras with involutions. J. Algebra, 130(1):65–82, 1990. [14] Max-Albert Knus and Manuel Ojanguren. Th´eorie de la descente et alg`ebres d’Azumaya. Lecture Notes in Mathematics, Vol. 389. Springer-Verlag, Berlin, 1974. [15] T. Y. Lam. Lectures on modules and rings, volume 189 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1999. [16] P. J. Morandi, B. A. Sethuraman, and J.-P. Tignol. Division algebras with an anti- automorphism but with no involution. Adv. Geom., 5(3):485–495, 2005. 28 RINGS THAT ARE MORITA EQUIVALENT TO THEIR OPPOSITES

[17] I. Reiner. Maximal orders, volume 28 of London Mathematical Society Monographs. New Series. The Clarendon Press Oxford University Press, Oxford, 2003. Corrected reprint of the 1975 original, With a foreword by M. J. Taylor. [18] Louis H. Rowen. Ring theory. Vol. I, volume 127 of Pure and Applied Mathematics. Academic Press Inc., Boston, MA, 1988. [19] David J. Saltman. Azumaya algebras with involution. J. Algebra, 52(2):526–539, 1978. [20] David J. Saltman. Lectures on division algebras, volume 94 of CBMS Regional Conference Series in Mathematics. Published by American Mathematical Society, Providence, RI, 1999. [21] Winfried Scharlau. Automorphisms and involutions of incidence algebras. In Proceedings of the International Conference on Representations of Algebras (Carleton Univ., Ottawa, Ont., 1974), Paper No. 24, pages 11 pp. Carleton Math. Lecture Notes, No. 9, Ottawa, Ont., 1974. Carleton Univ. [22] C. T. C. Wall. On the axiomatic foundations of the theory of Hermitian forms. Proc. Cam- bridge Philos. Soc., 67:243–250, 1970.